13.07.2015 Views

A review of porosity and Diffusion in Bentonite (pdf) (2.4 MB) - Posiva

A review of porosity and Diffusion in Bentonite (pdf) (2.4 MB) - Posiva

A review of porosity and Diffusion in Bentonite (pdf) (2.4 MB) - Posiva

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

Work<strong>in</strong>g Report 2013-29A Review <strong>of</strong> Porosity <strong>and</strong> <strong>Diffusion</strong><strong>in</strong> <strong>Bentonite</strong>C.A.J. AppeloHydrochemical ConsultantOctober 2013Work<strong>in</strong>g Reports conta<strong>in</strong> <strong>in</strong>formation on work <strong>in</strong> progressor pend<strong>in</strong>g completion.The conclusions <strong>and</strong> viewpo<strong>in</strong>ts presented <strong>in</strong> the reportare those <strong>of</strong> author(s) <strong>and</strong> do not necessarilyco<strong>in</strong>cide with those <strong>of</strong> <strong>Posiva</strong>.


1ABSTRACTPorosity <strong>in</strong> bentonite can be subdivided <strong>in</strong> free porewater, diffuse double layer (DDL)water, <strong>and</strong> <strong>in</strong>terlayer water <strong>in</strong> montmorillonite, the most important m<strong>in</strong>eral <strong>of</strong> bentonite.The proportions <strong>of</strong> these water types can be calculated from the <strong>in</strong>ternal <strong>and</strong> externalspecific surface area <strong>of</strong> montmorillonite <strong>and</strong> the width <strong>of</strong> <strong>in</strong>terlayer- <strong>and</strong> DDL-water.Stack<strong>in</strong>g <strong>of</strong> the TOT layers <strong>in</strong> montmorillonite <strong>in</strong> the c-direction is the master variablethat determ<strong>in</strong>es the specific <strong>in</strong>ternal <strong>and</strong> external surface areas. The thickness <strong>of</strong> the waterlayers depends on the ionic strength <strong>and</strong> the composition <strong>of</strong> the free porewater, <strong>and</strong>on the pack<strong>in</strong>g density <strong>of</strong> the bentonite.When the proportions <strong>of</strong> the water types change, the diffusion <strong>of</strong> cations, anions <strong>and</strong>neutral molecules is affected <strong>in</strong> different ways s<strong>in</strong>ce the diffusional properties vary. Diffusivefluxes are proportional with accessible <strong>porosity</strong>, chemical potential gradient, <strong>and</strong>diffusion coefficient, but not with concentration as has been proposed <strong>of</strong>ten <strong>in</strong> diffusionmodels. The equations for calculat<strong>in</strong>g the diffusive flux through <strong>in</strong>terlayer water are derived.The potential gradient here is given by the gradient <strong>of</strong> the equivalent (or molar)fraction <strong>of</strong> the cation <strong>in</strong> the cation exchange capacity. The latter is expressed as concentration<strong>in</strong> the <strong>in</strong>terlayer water. An example calculation illustrates that the flux <strong>in</strong> <strong>in</strong>terlayerwater can be dom<strong>in</strong>ant <strong>and</strong> opposite to the one <strong>in</strong> free porewater.Retardation by ion exchange is an important process <strong>in</strong> bentonite that can be modeled ifthe concentrations <strong>of</strong> major ions are known. Unfortunately, the analyses are almost neverdone <strong>in</strong> diffusion experiments.Keywords: <strong>Bentonite</strong>, montmorillonite, <strong>porosity</strong>, free porewater, diffuse double layer,<strong>in</strong>terlayer, specific surface area, water layer, diffusivity, diffusive flux.


2Katsaus bentoniit<strong>in</strong> huokoisuuteen ja diffuusio-om<strong>in</strong>aisuuksi<strong>in</strong>TIIVISTELMÄBentoniit<strong>in</strong> huokoisuus voidaan jaotella vapaaseen huokosveteen, diffusiivisessa kaksoiskerroksessaolevaan veteen (DDL) ja montmorilloniit<strong>in</strong> välikerroksissa olevaan veteen.Näiden vesityyppien suhteelliset osuudet voidaan laskea montmorilloniit<strong>in</strong> sisäistensekä ulkoisten om<strong>in</strong>aisp<strong>in</strong>ta-alojen ja väli- sekä DDL-kerrosten paksuuksista.Montmorilloniit<strong>in</strong> TOT-kerrosten p<strong>in</strong>outum<strong>in</strong>en c-suunnassa on keskeis<strong>in</strong> sisäiset jaulkoiset om<strong>in</strong>aisp<strong>in</strong>ta-alat määrittävä muuttuja. Vesikerrosten paksuus riippuu ionivahvuudestaja vapaan huokosveden koostumuksesta sekä bentoniit<strong>in</strong> pakkaustiheydestä.Eri vesityyppien suhteellisten osuuksien muuttuessa kationien, anionien ja neutraalienmolekyylien diffuusiot muuttuvat eri tavalla, koska niiden diffusiiviset om<strong>in</strong>aisuudetvaihtelevat. Diffusiiviset vuot ovat verrannollisia käytettävissä olevaan huokoisuuteen,kemiallisen potentiaal<strong>in</strong> gradientti<strong>in</strong> ja diffuusiokertoimeen muttei pitoisuuksi<strong>in</strong> kutenuse<strong>in</strong> on esitetty eri diffuusiomalleissa. Yhtälöt, joiden avulla lasketaan diffusiivisetvuot, on johdettu. Potentiaal<strong>in</strong> gradientti on tässä yhteydessä annettu kation<strong>in</strong>vaihtokapasiteet<strong>in</strong>kationien ekvivalent<strong>in</strong> (tai molaarisen) osuuden gradient<strong>in</strong> muodossa. Kation<strong>in</strong>vaihtokapasiteettiesitetään välikerroksen pitoisuutena. Esimerkilaskelman avullahava<strong>in</strong>nollistetaan välikerrosvedessä olevan vuon dom<strong>in</strong>oivan muih<strong>in</strong> nähden sekä senolevan vastakka<strong>in</strong>en vapaan huokosveden vuohon nähden.Ion<strong>in</strong> vaihdosta aiheutuva pidättym<strong>in</strong>en on merkittävä prosessi bentoniitissa. Tämä prosessivoidaan mall<strong>in</strong>taa mikäli merkittävimpien ionien pitoisuudet tunnetaan. Näitä analysejäei valitettavasti yleensä tehdä diffuusiokokeiden yhteydessä.Ava<strong>in</strong>sanat: Bentoniitti, montmorilloniitti, huokoisuus, vapaa huokosvesi, diffuusiv<strong>in</strong>enkaksoiskerros, välikerros, om<strong>in</strong>aisp<strong>in</strong>ta-ala, vesikerros, diffusiviteetti, diffusiiv<strong>in</strong>envuo.


1TABLE OF CONTENTSABSTRACTTIIVISTELMÄ1 INTRODUCTION ................................................................................................... 22 POROSITY IN BENTONITE .................................................................................. 42.1 A structural model for <strong>porosity</strong> <strong>in</strong> bentonite .................................................. 42.2 Distribut<strong>in</strong>g the water ................................................................................... 52.3 Calculat<strong>in</strong>g the porewater distribution from experimental data .................... 73 DIFFUSION IN BENTONITE: INTRODUCTION ................................................. 123.1 Models for diffusion <strong>in</strong> bentonite ................................................................ 163.2 Which concentration gradient should be used? ......................................... 193.3 Example: fluxes <strong>of</strong> I - , Cs + <strong>and</strong> Sr 2+ <strong>in</strong> the 3 <strong>porosity</strong> doma<strong>in</strong>s ................... 204 DISCUSSION ...................................................................................................... 234.1 Experimental verification is necessary; consistency checks are possible .. 255 CONCLUSIONS .................................................................................................. 276 ACKNOWLEDGMENTS ...................................................................................... 28REFERENCES ............................................................................................................. 29


21 INTRODUCTION<strong>Bentonite</strong> is foreseen as a conf<strong>in</strong><strong>in</strong>g material <strong>in</strong> the disposal <strong>of</strong> radioactive waste (<strong>Posiva</strong>2008; NAGRA 2002). <strong>Bentonite</strong> conta<strong>in</strong>s a high proportion <strong>of</strong> Na-montmorillonite, aclay m<strong>in</strong>eral that swells <strong>in</strong> water <strong>and</strong> endows bentonite with a very low hydraulic conductivity(Pusch, 2001). As a result, transport <strong>of</strong> water <strong>and</strong> solutes <strong>in</strong> bentonite is ma<strong>in</strong>lydiffusive, with diffusion coefficients that are much smaller than <strong>in</strong> ord<strong>in</strong>ary water (Madsen1998; Ochs et al. 2001; Bourg et al. 2007). Moreover, montmorillonite carries anegative charge that is balanced by cations which exchange readily with other solutecations (Sposito 1984). Cation exchange retards solute transport (Appelo <strong>and</strong> Postma2005), <strong>and</strong> further enhances the barrier properties <strong>of</strong> bentonite that stem from the negligibleadvective flow <strong>and</strong> the small diffusion coefficients. However, for the performance<strong>of</strong> a repository, even small transport rates <strong>of</strong> radioactive elements with extended halflivescan be important. Moreover, <strong>in</strong>teraction with the environment can change theproperties <strong>of</strong> bentonite. For example, Ca 2+ from external water can exchange with Na +by diffusion <strong>and</strong> cause the montmorillonite structure to collapse (Laird 2006). As a result,the hydraulic conductivity may <strong>in</strong>crease markedly. Chemical calculations <strong>of</strong> these<strong>in</strong>teractions are done with geochemical models that can <strong>in</strong>corporate solubility limitationson the concentrations <strong>of</strong> various elements, as well as diffusion. However, the modelsare not well adapted to h<strong>and</strong>le the specific properties <strong>of</strong> bentonite, <strong>in</strong> which water isdistributed over diverse porosities with different characteristics.This report gives a <strong>review</strong> <strong>of</strong> the <strong>porosity</strong> structure <strong>of</strong> bentonite <strong>and</strong> diffusion modelsadopted so far. The <strong>in</strong>ner structure <strong>of</strong> bentonite with its different porosities is <strong>of</strong> paramountimportance for underst<strong>and</strong><strong>in</strong>g how transport (by diffusion) depends on compaction<strong>and</strong> solution composition. A model structure is derived that allows to calculate the<strong>porosity</strong> distribution as a function <strong>of</strong> the layer stack<strong>in</strong>g <strong>of</strong> montmorillonite <strong>and</strong> the NaClconcentration <strong>in</strong> free porewater. The model is used to extract the <strong>porosity</strong> distributionfrom experiments <strong>in</strong> which the accessible <strong>porosity</strong> <strong>of</strong> Cl - was measured as a function <strong>of</strong>pack<strong>in</strong>g density <strong>and</strong> NaCl concentration (Muur<strong>in</strong>en 2006; Van Loon et al. 2007). Themodel structure can be <strong>in</strong>troduced <strong>in</strong> general geochemical transport models. <strong>Diffusion</strong>models are <strong>review</strong>ed, <strong>and</strong> a model is derived for diffusion <strong>of</strong> exchangeable cations <strong>in</strong><strong>in</strong>terlayer water.


3Figure 1. Illustrations <strong>of</strong> the bentonite structure <strong>and</strong> <strong>porosity</strong> types.(A): A scann<strong>in</strong>g electron micrograph <strong>of</strong> packed bentonite <strong>in</strong>dicat<strong>in</strong>g <strong>in</strong>tergranular <strong>porosity</strong><strong>and</strong> macroscopic layer<strong>in</strong>g as a result <strong>of</strong> compaction. The <strong>in</strong>tergranular <strong>porosity</strong>is filled with free porewater. The scale bar is 100 μm (Sato <strong>and</strong> Suzuki 2003).(B): A high-resolution transmission electron micrograph show<strong>in</strong>g the 1 nm thick TOTlayers<strong>of</strong> a montmorillonite quasicrystal. Interlayer water resides between the layers.The scale bar is 10 nm (Laird et al. 1989).(C): <strong>Bentonite</strong> with montmorillonite <strong>and</strong> m<strong>in</strong>eral impurities (these may have larger crystalsize than <strong>in</strong>dicated, Mäder, pers. comm.) illustrat<strong>in</strong>g the location <strong>of</strong> differentwater types (Bradbury <strong>and</strong> Baeyens 2002).(D): A sketch <strong>of</strong> water <strong>in</strong> montmorillonite show<strong>in</strong>g the heterogeneous distribution <strong>of</strong>anions <strong>and</strong> cations <strong>in</strong> three water types. Interlayer water has cations only, diffusedouble layer water has an excess <strong>of</strong> cations, <strong>and</strong> free porewater is charge-neutral.TOT <strong>in</strong>dicates the tetrahedral-octahedral-tetrahedral coord<strong>in</strong>ation <strong>of</strong> oxygenaround cations <strong>in</strong> the structure (Wers<strong>in</strong> et al. 2004)


42 POROSITY IN BENTONITELaboratory experiments have shown that tracer diffusion coefficients change with thepreparation <strong>and</strong> pack<strong>in</strong>g density <strong>of</strong> the bentonite <strong>and</strong> with the concentrations <strong>of</strong> the majorions (Sato et al. 1992; Kim et al. 1993; Muur<strong>in</strong>en 1994, 2006; Kozaki et al. 1998a, b,2001, 2008; Molera <strong>and</strong> Eriksen 2002; Oscarson 1994; Van Loon et al. 2007, García-Gutiérrez et al. 2004). The changes are related to the <strong>in</strong>ner structure <strong>of</strong> a compactedbentonite with different porosities as illustrated <strong>in</strong> Figure 1. Thus, when calculat<strong>in</strong>g diffusionalfluxes <strong>in</strong> bentonite, the first challenge is to f<strong>in</strong>d how porewater is distributedover these porosities. Three types <strong>of</strong> water with different properties emerge from Figure1 (Kozaki et al. 2008; Bradbury <strong>and</strong> Baeyens 2002; Bourg et al. 2003; Wers<strong>in</strong> et al.2004; Muur<strong>in</strong>en 2006):Interlayer water (ε IL ) with water <strong>and</strong> cations between the TOT layers <strong>of</strong>montmorillonite, devoid <strong>of</strong> anions. The cations compensate the structural charge deficit<strong>of</strong> the TOT layers, the water molecules are built up <strong>in</strong> layers <strong>and</strong> are part <strong>of</strong> the crystallographicstructure <strong>of</strong> the montmorillonite.Diffuse double layer (DDL) water (ε DDL ), conta<strong>in</strong><strong>in</strong>g water, cations <strong>and</strong> anions; an excess<strong>of</strong> cations neutralizes the charge at the outer surface <strong>of</strong> the montmorillonite. TheDDL forms the transition zone from the m<strong>in</strong>eral surface to free porewater.Free porewater (ε free ), a charge-balanced aqueous solution <strong>of</strong> cations <strong>and</strong> anions.The proportions <strong>of</strong> the water types depend on size <strong>and</strong> shape <strong>of</strong> the montmorilloniteflakes <strong>in</strong> the bentonite <strong>and</strong> the degree <strong>of</strong> compaction expressed by the dry density. Furthermore,the free porewater composition determ<strong>in</strong>es, by electrostatics, the extent <strong>of</strong> thediffuse double layer <strong>in</strong> the pore <strong>and</strong> <strong>in</strong>fluences, by osmotic pressure, the <strong>in</strong>terlayer water(Kozaki et al. 2008).2.1 A structural model for <strong>porosity</strong> <strong>in</strong> bentoniteThe effect <strong>of</strong> size <strong>and</strong> shape <strong>of</strong> montmorillonite on the <strong>porosity</strong> can be calculated bystack<strong>in</strong>g a montmorillonite unit-cell <strong>in</strong> 3 dimensions <strong>and</strong> calculat<strong>in</strong>g the <strong>in</strong>ternal <strong>and</strong>external surface. The chemical formula <strong>of</strong> a typical montmorillonite isNa 0.6 [Si 7.8 Al 0.2 ] IV [Al 3.6 Mg 0.4 ] VI O 20 (OH) 4 (omitt<strong>in</strong>g water), with molecular weight MW =0.735 kg/mol <strong>and</strong> crystal density ρ cryst = 2760 kg/m 3 . The chemical formula is for a unitcellwith the dimensions a × b × c * = 0.523 × 0.905 × 0.96 nm 3 , where c * = c s<strong>in</strong>(95°) isthe orthogonally projected c-axis <strong>of</strong> the monocl<strong>in</strong>ic unit-cell. When montmorilloniteexp<strong>and</strong>s <strong>in</strong> water, c * <strong>in</strong>creases with t IL , the <strong>in</strong>terlayer thickness which is variable <strong>and</strong> dependson the number <strong>of</strong> water molecules adsorbed. The unit-cells are stacked <strong>in</strong> 3 dimensionsby n a , n b <strong>and</strong> n c .The external surface area is:Aext**na nca c nb ncb c na nba b N A 2m/kg 2(1)n n nMWwhere N A = 6.022×10 23 molecules/mol (Avogadro’s number).abc


5The <strong>in</strong>ternal planar surface is:Aplanar,<strong>in</strong>tnc1 N A 2m/kga b 2(2)n MWcThe average diameter <strong>of</strong> montmorillonite flakes is about 0.15 μm (Pusch 2001) whichsuggests that n a <strong>and</strong> n b are close to 200. An edge surface <strong>of</strong> 3 m 2 /g is <strong>of</strong>ten used for calculat<strong>in</strong>gproton buffer<strong>in</strong>g <strong>and</strong> specific sorption with the surface complexation model. Ifthe surface complexation properties are assigned to the octahedral (O) layer, which is0.4 nm thick, the edge surface area <strong>of</strong> 3 m 2 /g results when n a = n b = 311, assum<strong>in</strong>g thatstack<strong>in</strong>g is the same along a <strong>and</strong> b. The value <strong>of</strong> n c ranges from 3 – 5 for Namontmorillonite,<strong>and</strong> is about 10 when divalent cations occupy the <strong>in</strong>terlayer space(Pusch 2001). With such small n c , <strong>and</strong> high n a <strong>and</strong> n b , the contribution <strong>of</strong> the edge surfacearea to the total external surface is relatively small <strong>and</strong> not much dependent on theexact values <strong>of</strong> n a <strong>and</strong> n b .2.2 Distribut<strong>in</strong>g the waterThe sum <strong>of</strong> <strong>in</strong>terlayer-, DDL- <strong>and</strong> free water-<strong>porosity</strong> (ε tot ) is given by:drytot1 (3)crystwhere ρ dry is the dry density <strong>of</strong> the packed bentonite (kg/m 3 ).The <strong>in</strong>terlayer <strong>porosity</strong> can be calculated multiply<strong>in</strong>g the <strong>in</strong>ternal planar surface with the<strong>in</strong>terlayer thickness <strong>and</strong> the relative density <strong>of</strong> the <strong>in</strong>terlayer water. The <strong>in</strong>terlayer thicknessis a variable which depends on the type <strong>of</strong> montmorillonite (Slade et al., 1991), theactivity <strong>of</strong> water <strong>and</strong> the type <strong>of</strong> cation (Cases et al. 1992, 1997; Bérend et al. 1995;Ferrage et al. 2005b) <strong>and</strong> on the compaction <strong>of</strong> the bentonite expressed by the dry density(Kozaki 1998a; Muur<strong>in</strong>en 2007). For Na + , three dist<strong>in</strong>ct spac<strong>in</strong>gs can be observed <strong>in</strong>XRD-patterns for 1-, 2- <strong>and</strong> 3-layer hydrates, with gradual transitions <strong>in</strong>-between(Ferrage et al. 2005b; Slade et al. 1991). For bentonite packed at dry densities vary<strong>in</strong>gfrom 1000 to 1800 kg/m 3 , Kozaki noted d 001 spac<strong>in</strong>gs <strong>of</strong> 1.88 <strong>and</strong> 1.56 nm, correspond<strong>in</strong>gto 3- <strong>and</strong> 2-layer hydrates, respectively, with the transition occurr<strong>in</strong>g between 1300<strong>and</strong> 1600 kg/m 3 . Based on these, <strong>and</strong> his own data, Muur<strong>in</strong>en et al. (2007) proposed acont<strong>in</strong>uous decrease <strong>of</strong> the <strong>in</strong>terlayer thickness <strong>in</strong> the range from 750 to 1800 kg/m 3 accord<strong>in</strong>gto:t IL = 1.41×10 -9 – 4.9×10 -13 ρ dry (m) (4)Bourg et al. (2006), also build<strong>in</strong>g on the Kozaki et al. data, positioned the transition <strong>of</strong>the three- <strong>and</strong> two-layer hydrates <strong>in</strong>-between 1420 <strong>and</strong> 1570 kg/m 3 .The <strong>in</strong>terlayer space contracts with <strong>in</strong>creas<strong>in</strong>g NaCl concentration (Norrish <strong>and</strong> Quirk1954; Slade et al. 1991; Kozaki et al. 2008), <strong>and</strong> an additional term can be added to accountfor this osmotic effect:


6t IL = 1.41×10 -9 – 4.9×10 -13 ρ dry + f I_IL × (m ext – 0.2) (m) (5)where (m ext – 0.2) is the total concentration <strong>of</strong> solutes <strong>in</strong> free porewater (mol/L) with 0.2M as reference level (Kozaki et al. (2008) observed the osmotic contraction to start at0.1 M NaCl), <strong>and</strong> f I_IL is a coefficient (expected to be negative). The osmotic effect is, <strong>in</strong>pr<strong>in</strong>ciple, l<strong>in</strong>ear with concentration. This contrasts with the expansion <strong>of</strong> a s<strong>in</strong>gle, unconf<strong>in</strong>edflake <strong>of</strong> montmorillonite with (c NaCl ) -0.5 , which <strong>in</strong>dicates an <strong>in</strong>crease <strong>of</strong> theDDL thickness (Norrish 1954).The volume <strong>of</strong> the <strong>in</strong>terlayer space isV IL = t IL × A planar,<strong>in</strong>t / 2 (m 3 /kg) (6)The <strong>in</strong>terlayer <strong>porosity</strong> becomes:ε IL = V IL × f dens × w mm × ρ dry (-) (7)where f dens is the density ratio <strong>of</strong> water <strong>in</strong> the <strong>in</strong>terlayer space <strong>and</strong> <strong>in</strong> the free pore (-),<strong>and</strong> w mm is the weight fraction <strong>of</strong> montmorillonite <strong>in</strong> the bentonite. The value <strong>of</strong> f dens isnot exactly known, as discussed <strong>in</strong> Section 5.Assum<strong>in</strong>g for a start that f dens = 1, <strong>and</strong> f I_IL = 0, Figure 2 presents the <strong>in</strong>terlayer <strong>porosity</strong>for n c = 1.5, 3 <strong>and</strong> 25. Also, the <strong>in</strong>terlayer <strong>porosity</strong> is drawn accord<strong>in</strong>g to Bourg et al.(2006) who change the <strong>in</strong>terlayer thickness from the 2- <strong>in</strong>to the 3-layer hydrate between1420 <strong>and</strong> 1570 kg/m 3 <strong>and</strong> assume that all the basal planes are border<strong>in</strong>g on <strong>in</strong>terlayerwater (n c = ∞). F<strong>in</strong>ally, a simple model shows <strong>in</strong>terlayer water as a constant fraction <strong>of</strong>the dry density <strong>of</strong> 0.16×10 -3 / (kg/m 3 ) (Bourg et al. 2003).Figure 2. Interlayer <strong>porosity</strong> <strong>in</strong> Na-montmorillonite as a function <strong>of</strong> the dry density accord<strong>in</strong>gto various models:


71) the <strong>in</strong>terlayer thickness decreases gradually with the dry density (Muur<strong>in</strong>en et al.,2007); the <strong>in</strong>terlayer <strong>porosity</strong> depends on unit-cell stack<strong>in</strong>g <strong>in</strong> the c-direction (n c ) <strong>and</strong>is shown for n c = 1.5, 3 <strong>and</strong> 25 (l<strong>in</strong>es);2) the <strong>in</strong>terlayer thickness changes from 0.9 to 0.55 nm between dry densities <strong>of</strong> 1420<strong>and</strong> 1570 kg/m 3 <strong>and</strong> all the planar surfaces are border<strong>in</strong>g the <strong>in</strong>terlayer space (n c = ∞;Bourg et al. 2006, dotted l<strong>in</strong>e);3) the <strong>in</strong>terlayer <strong>porosity</strong> is a constant fraction <strong>of</strong> the dry density <strong>of</strong> 0.16 (Bourg et al.2003, dashed l<strong>in</strong>e).First, figure 2 illustrates that the <strong>in</strong>terlayer <strong>porosity</strong> is a very significant part <strong>of</strong> the total<strong>porosity</strong> when dry density <strong>in</strong>creases above 1000 kg/m 3 . This is reached <strong>in</strong> all practicalapplications <strong>of</strong> bentonite. At a dry density <strong>of</strong> 1800 kg/m 3 almost all the <strong>porosity</strong> resides<strong>in</strong> the <strong>in</strong>terlayers <strong>of</strong> Na-montmorillonite. S<strong>in</strong>ce anions are excluded from the <strong>in</strong>terlayers(Bolt <strong>and</strong> De Haan, 1982), the anion-accessible <strong>porosity</strong> becomes zero, <strong>and</strong> aniondiffusionis m<strong>in</strong>imized (Bourg et al. 2003).Second, the number <strong>of</strong> stacks <strong>in</strong> the c-direction has considerable <strong>in</strong>fluence on the <strong>in</strong>terlayer<strong>porosity</strong>. The <strong>in</strong>terlayer <strong>porosity</strong> <strong>in</strong>creases with n c . The maximum is reached whenn c ≈ 25 <strong>and</strong> it rema<strong>in</strong>s almost the same with higher n c . The <strong>in</strong>terlayer <strong>porosity</strong> halveswith n c when n c is smaller than 3, <strong>and</strong> it becomes zero for n c = 1. Thus, a well washed<strong>and</strong> completely dispersed Na-montmorillonite <strong>in</strong> distilled water will have zero <strong>in</strong>terlayer<strong>porosity</strong>.The DDL <strong>porosity</strong> can be calculated by multiply<strong>in</strong>g the external surface area with thethickness <strong>of</strong> the DDL. The thickness <strong>of</strong> the DDL can be expressed <strong>in</strong> Debye-lengths:t10DDL f DDL3.0910 m(8)IWhere I is the ionic strength (mol/L), <strong>and</strong> f DDL is the number <strong>of</strong> Debye-lengths. TheDDL <strong>porosity</strong> is:ε DDL = A ext × t DDL × w mm × ρ dry (-) (9)Def<strong>in</strong>ed <strong>in</strong> this way, f DDL also corrects the decrease <strong>of</strong> external surface area by gra<strong>in</strong> togra<strong>in</strong> contacts.F<strong>in</strong>ally, the free water <strong>porosity</strong> becomes:ε free = ε tot - ε IL - ε DDL (-) (10)2.3 Calculat<strong>in</strong>g the porewater distribution from experimental dataMuur<strong>in</strong>en (2006) <strong>and</strong> Van Loon et al. (2007) have measured the anion accessible <strong>porosity</strong><strong>of</strong> bentonite as a function <strong>of</strong> the dry density <strong>and</strong> the concentration <strong>of</strong> NaCl <strong>in</strong> theexternal solution. In pr<strong>in</strong>ciple, the experimental data permit to obta<strong>in</strong> the coefficients n c ,f I_IL , f dens <strong>and</strong> f DDL . However, the coefficients were found to be highly correlated <strong>and</strong> the


8data are few <strong>and</strong> not very precise (cf. Van Loon et al. 2007). Thus, the model was simplifiedto derive the effect <strong>of</strong> the NaCl-concentration for a given bentonite dry density.Accord<strong>in</strong>g to Equations (5) <strong>and</strong> (8), the external solution concentration affects the anionaccessible <strong>porosity</strong> <strong>in</strong> two ways. The <strong>in</strong>terlayer <strong>porosity</strong> decreases l<strong>in</strong>early with concentrationby the osmotic effect, <strong>and</strong> the DDL <strong>porosity</strong> decreases with the square root <strong>of</strong>concentration. Or, <strong>in</strong> Equation (10), assum<strong>in</strong>g that DDL-water <strong>and</strong> <strong>in</strong>terlayer water aredevoid <strong>of</strong> anions:aDDL free tot a0 _ IL aosm I NaCl (11)INaClwhere a 0_IL , a osm <strong>and</strong> a DDL are variables that lump the effects <strong>of</strong> the ‘f’’ parameters fromthe previous section.The parameters were optimized to obta<strong>in</strong> the smallest sum <strong>of</strong> the squared deviations (us<strong>in</strong>gPEST, Doherty 1994) <strong>and</strong> the result<strong>in</strong>g Cl - -accessible porosities (assumed equal toε free ) are shown <strong>in</strong> Figures 3 <strong>and</strong> 4, together with the data. Muur<strong>in</strong>en used MX80 <strong>and</strong>washed it <strong>in</strong> NaCl, while Van Loon et al. used Volclay as received from the manufacturer.The Cl - -accessible porosities were measured for different dry densities <strong>of</strong> 1300,1600, <strong>and</strong> 1700 or 1900 kg/m 3 , <strong>and</strong> different concentrations <strong>of</strong> NaCl <strong>in</strong> the external solution.At the same ionic strength <strong>and</strong> dry density, Van Loon et al. found a smaller Cl - -accessible <strong>porosity</strong> than Muur<strong>in</strong>en, possibly a result <strong>of</strong> the different orig<strong>in</strong> <strong>of</strong> thebentonite <strong>and</strong> the different preparation. However, both datasets show the expected trend<strong>of</strong> a decreas<strong>in</strong>g Cl - -accessible <strong>porosity</strong> with <strong>in</strong>creas<strong>in</strong>g dry density <strong>of</strong> the bentonite <strong>and</strong>with decreas<strong>in</strong>g NaCl concentration, <strong>and</strong> the data are closely matched by the modell<strong>in</strong>es.Parts B, C <strong>and</strong> D <strong>in</strong> figures 3 <strong>and</strong> 4 illustrate the model’s distribution <strong>of</strong> the <strong>porosity</strong>over free porewater, <strong>in</strong>terlayer <strong>and</strong> DDL- water, where the latter two are given by the2 nd <strong>and</strong> 3 rd terms <strong>of</strong> the right-h<strong>and</strong> side <strong>of</strong> Equation (11). In both datasets, the fraction <strong>of</strong><strong>in</strong>terlayer water <strong>in</strong>creases, as expected, with decreas<strong>in</strong>g NaCl concentration <strong>and</strong> with<strong>in</strong>creas<strong>in</strong>g dry density. The fraction <strong>of</strong> DDL-water is substantial <strong>in</strong> Muur<strong>in</strong>en’sbentonite, but fairly small <strong>in</strong> the bentonite used by Van Loon et al. For NaCl concentrationsbelow 0.4 M, the fraction <strong>of</strong> DDL <strong>porosity</strong> <strong>in</strong>creases markedly with decreas<strong>in</strong>gNaCl concentration. Concomitantly, the anion-accessible <strong>porosity</strong> dim<strong>in</strong>ishes to verysmall values.


9Figure 3. (A): Anion-accessible <strong>porosity</strong> <strong>in</strong> bentonite packed at different dry density asa function <strong>of</strong> NaCl concentration <strong>in</strong> the external solution, <strong>and</strong> (B, C, D): modeldistribution<strong>of</strong> water over Cl - -accessible (assumed equal to ε free ), DDL- (ε DDL ) <strong>and</strong> <strong>in</strong>terlayer(ε IL ) porosities for pack<strong>in</strong>gs <strong>of</strong> (B) 1280, (C) 1580, (D) 1700 kg/m 3 . Experimentaldata from Muur<strong>in</strong>en (2006).The optimized values <strong>of</strong> the parameters a 0_IL , a osm <strong>and</strong> a DDL are listed <strong>in</strong> Table 1. Thenumbers express the different trends <strong>in</strong> the two datasets. In Van Loon’s experiments, theCl - -accessible <strong>porosity</strong> is dom<strong>in</strong>ated by the change <strong>of</strong> the <strong>in</strong>terlayer <strong>porosity</strong> (relativelylarge numbers for a 0_IL <strong>and</strong> a osm ), whereas the DDL <strong>porosity</strong> is <strong>in</strong>fluential only whenionic strength decreases below 0.05. Muur<strong>in</strong>en’s data show equal importance <strong>of</strong> both<strong>porosity</strong> types, although also here, the DDL-<strong>porosity</strong> becomes predom<strong>in</strong>ant when theionic strength becomes less than 0.1.The numbers permit calculation <strong>of</strong> the specific external <strong>and</strong> <strong>in</strong>ternal surface area <strong>of</strong> themontmorillonite. The external surface area is:Aext3.09 1010afDDLDDLwmmdryIt was assumed that f DDL = 1, <strong>and</strong> w mm = 0.8 <strong>and</strong> 0.71 for Muur<strong>in</strong>en <strong>and</strong> Van Loon’s data,respectively.(12)


10Figure 4. (A): Anion-accessible <strong>porosity</strong> <strong>in</strong> bentonite packed at different dry density asa function <strong>of</strong> NaCl concentration <strong>in</strong> the external solution, <strong>and</strong> (B, C, D): modeldistribution<strong>of</strong> water over Cl - -accessible (ε free ), DDL- (ε DDL ) <strong>and</strong> <strong>in</strong>terlayer (ε IL ) porositiesfor pack<strong>in</strong>gs <strong>of</strong> (B) 1300, (C) 1600, (D) 1900 kg/m 3 . The DDL-<strong>porosity</strong> is represented<strong>in</strong> B-D by the small field <strong>in</strong>-between the dashed l<strong>in</strong>e <strong>and</strong> the full l<strong>in</strong>e that follows thedata. Experimental data from Van Loon et al. (2007).The specific <strong>in</strong>terlayer surface area is:Aplanar ,<strong>in</strong>t2 a0 _ IL a-9-131.4110- 4.910dry+ fI_ IL (2INaCl - 0.2) fdenswmmdryosm INaCl(13)It was assumed that f dens = 1, <strong>and</strong> the numbers <strong>in</strong> Table 1 are for I NaCl = 0.1.The specific surface areas for Muur<strong>in</strong>en’s data are <strong>in</strong> good agreement with expectedvalues for montmorillonite, but the <strong>in</strong>ternal planar surface area is too high for VanLoon’s data. Of course, this is related to the different fit-parameters (Table 1). The fitparametersexpress the different trends <strong>of</strong> the Cl - -accessible <strong>porosity</strong> that can be noted<strong>in</strong> Figures 3A <strong>and</strong> 4A. In Van Loon’s experiments, the Cl-accessible <strong>porosity</strong> <strong>in</strong>creases


11l<strong>in</strong>early with the NaCl concentration, <strong>in</strong> Muur<strong>in</strong>en’s data the <strong>in</strong>crease flattens out. Thus,<strong>in</strong> Van Loon’s data the osmotic component is more important than <strong>in</strong> Muur<strong>in</strong>en’s, <strong>in</strong>which the DDL parameter predom<strong>in</strong>ates (Table 1). It may be that the preparation, or rather,the lack <strong>of</strong> preparation <strong>of</strong> the bentonite bears upon these results. The Volclay usedby Van Loon conta<strong>in</strong>s calcite <strong>and</strong> perhaps gypsum, which dissolve more when the ionicstrength <strong>in</strong>creases. The <strong>in</strong>creas<strong>in</strong>g Ca 2+ concentration may cause the <strong>in</strong>terlayer space tocollapse, thus caus<strong>in</strong>g the free <strong>porosity</strong> to <strong>in</strong>crease with ionic strength (Dufey et al.1976). Muur<strong>in</strong>en washed MX80 bentonite with NaCl solution <strong>and</strong> may have removedthe more readily dissolv<strong>in</strong>g m<strong>in</strong>erals. Muur<strong>in</strong>en (2007) also experimented withbentonite that was washed with distilled water, <strong>and</strong> observed that the Cl - -accessible porositieswere smaller than without wash<strong>in</strong>g.Table 1. Fit-parameters for the Cl - -accessible <strong>porosity</strong> measured by Muur<strong>in</strong>en (2006)<strong>and</strong> Van Loon et al. (2007) as a function <strong>of</strong> the dry density <strong>of</strong> bentonite. The fitparametersallow to obta<strong>in</strong> the <strong>in</strong>ternal planar, <strong>and</strong> total external surface <strong>of</strong> themontmorillonite, <strong>and</strong> to estimate the stack<strong>in</strong>g number <strong>in</strong> the c direction.Muur<strong>in</strong>en(2006)Van Loon etal. (2007)ρ dry(kg/m 3 )a 0_IL(-)a osm(mol/L) -1a DDL(mol/L) -0.5A planar,<strong>in</strong>t(m 2 /g)A ext(m 2 /g)1280 ± 80 0.162 -0.0106 0.0749 402 2371580 ± 40 0.192 -0.00856 0.0575 476 1471700 ± 70 0.299 -0.0128 0.0363 583 861300 0.478 -0.321 0.00312 1250 111600 0.388 -0.0632 0.00291 1073 81900 0.300 -0.0244 0.00157 921 4n c2.6>100


123 DIFFUSION IN BENTONITE: INTRODUCTIONThe diffusional flux <strong>of</strong> species i is given by:Jiuicii | z | F xi(14)where J i is the flux (mol/m 2 /s), u i is the mobility (m 2 /s/V), c i is the concentration(mol/m 3 ), z i is charge number (-), F is the Faraday constant (96485 J/V/eq), x is distance(m), <strong>and</strong> μ i is the thermodynamic potential, given by:μ i = μ i 0 + RT ln a i + z i Fψ (15)where μ i 0 is the st<strong>and</strong>ard potential (J/mol), R is the gas constant (8.314 J/K/mol), T isthe absolute temperature (K), a i is the activity (-), <strong>and</strong> ψ is the electrical potential (V).The activity is related to concentration by a i = γ i c i /c 0 , where γ i is the activity coefficient(-) <strong>and</strong> c i 0 is the st<strong>and</strong>ard state (1 mol/kg H 2 O, assumed equal to 1 mol/L <strong>in</strong> the follow<strong>in</strong>g).Assum<strong>in</strong>g an activity coefficient γ i = 1 <strong>and</strong> an electric potential gradient ∂ψ/∂x = 0, <strong>and</strong>us<strong>in</strong>g the identities u i = D w, i |z i | F / (RT) <strong>and</strong> c d(ln c) = d(c), Equation (14) becomes:Jidci Dw,i(16)dxwhere D w,i is the diffusion coefficient <strong>in</strong> water (m 2 /s) <strong>and</strong> c i has units mol/m 3 . In a porousmedium, the porewater diffusion coefficient differs from the diffusion coefficient<strong>in</strong> free water (D w,i ) by tortuosity <strong>and</strong> constrictivity factors (Van Brakel <strong>and</strong> Heertjes,1974; Figure 5):D p ,i D2 w,i(17)The tortuosity factor (θ 2 ) expresses that diffus<strong>in</strong>g molecules have to pass around solidgra<strong>in</strong>s <strong>and</strong> take a longer path (L a ) than the straight l<strong>in</strong>e distance (L) (Figure 5). A tortuouspath at an angle <strong>of</strong> 45° with the straight l<strong>in</strong>e distance obviously gives a tortuosity θ= 2 / √2 = 1.4, <strong>and</strong> a tortuosity factor θ 2 = 2. The constrictivity factor (δ) encompasseseffects <strong>of</strong> pore narrow<strong>in</strong>g <strong>and</strong> widen<strong>in</strong>g. A straight pore has a constrictivity factor <strong>of</strong> 1,if the pore narrows it becomes smaller than 1, <strong>and</strong> if it widens, larger than 1. Inert porousmedia have constrictivity factors slightly smaller than 1 (Van Brakel <strong>and</strong> Heertjes1974). However, <strong>in</strong> bentonite where DDL water occupies part <strong>of</strong> the pore space, porenarrow<strong>in</strong>g that results <strong>in</strong> overlapp<strong>in</strong>g DDL’s has a different effect for cations <strong>and</strong> neutralspecies that still can pass through, <strong>and</strong> for anions for which the constriction becomesan obstruction that has to be circumnavigated. In that case the constrictivity issmaller for anions than for cations <strong>and</strong> neutral species, while at the same time, for anionsthe tortuosity <strong>in</strong>creases.


13Figure 5. (Left:) The tortuosity is the ratio <strong>of</strong> traveled <strong>and</strong> straight-l<strong>in</strong>e distance,squared <strong>in</strong> the tortuosity factor (Appelo <strong>and</strong> Postma, 2005). (Right:) Overall picture <strong>of</strong>the <strong>porosity</strong> <strong>in</strong> clay, show<strong>in</strong>g changes <strong>in</strong> pore-width (constrictivity). Overlapp<strong>in</strong>g doublelayers at pore constrictions <strong>in</strong>crease the tortuosity for anions (Appelo et al. 2008).The quotient <strong>of</strong> the constrictivity <strong>and</strong> tortuosity factors is assembled <strong>in</strong> the geometricalfactor (G i ) which is related by Archie law to the accessible <strong>porosity</strong> (ε a,i ) (Van Loon etal., 2007; Grathwohl 1998): 12Gi na,i(18)where n is an empirical factor (≈ 1).In the flux equation, the porewater diffusion coefficient <strong>of</strong> the porous medium (D p,i ) iscorrected for the surface area that is accessible for the solute. Thus:D e, i = D(19)a,ip, iComb<strong>in</strong><strong>in</strong>g Equations (16-19) gives the flux <strong>in</strong> terms <strong>of</strong> the known diffusion coefficient<strong>in</strong> free water <strong>and</strong> accessible <strong>porosity</strong> as:Ji1 n dci a,iDw,i(20)dxOften <strong>in</strong> experiments, the concentration pr<strong>of</strong>ile <strong>in</strong> the bentonite is measured <strong>and</strong> <strong>in</strong>terpreted<strong>in</strong> terms <strong>of</strong> an apparent diffusion coefficient:citDa,i2ci2x(21)where D a,i is the apparent diffusion coefficient (m 2 /s), <strong>and</strong> where c i can be the concentration<strong>in</strong> porewater as before, or the concentration <strong>in</strong> the bulk volume (= ε a, i c i ) asmeasured easily when i is a radioactive tracer. The apparent diffusion coefficient <strong>in</strong>cludesthe effects <strong>of</strong> retardation by sorption, ion-exchange, <strong>and</strong>/or precipitation <strong>and</strong> dissolution:


14Dp,iDa,i (22)Rwhere R is the retardation (-), given by:Rddsi 1(23) dca,iiwhere ρ d is the dry density <strong>of</strong> solid (kg/m 3 ) <strong>and</strong> s i the concentration <strong>in</strong> the solid(mol/kg). Usually for tracers, the slope <strong>of</strong> the isotherm (ds i / dc i ) is constant <strong>and</strong> can beestimated. For Na-montmorillonite it is (Appelo <strong>and</strong> Postma 2005):z idsiCEC KI\KdcizNa i a(24)d 1000 Na where K d is the distribution coefficient (m 3 /kg), CEC is the cation exchange capacity (≈1 eq/kg for montmorillonite), K I\Na is the exchange constant (-), <strong>and</strong> a Na is the activity <strong>of</strong>Na + <strong>in</strong> free porewater, <strong>in</strong> the first <strong>in</strong>stance about equal to the NaCl concentration. Values<strong>of</strong> K I\Na are given by Appelo <strong>and</strong> Postma (2005).Altogether, Equations (21-24) relate the chang<strong>in</strong>g concentration <strong>of</strong> a tracer with time<strong>and</strong> the porewater diffusion coefficient as:cita,ia,iDdp,iKd2ci2x(25)The apparent diffusion coefficient <strong>in</strong> Na-montmorillonite can be found from:Da,ia,i d1na,i10z3iDw,i K aI\ NaNaz i(26)Equation (26) conta<strong>in</strong>s terms that should be known for the experiment or can be lookedup, <strong>and</strong> is useful for obta<strong>in</strong><strong>in</strong>g first estimates.Figure 6 compares measured apparent diffusion coefficients <strong>of</strong> Na + , Cs + <strong>and</strong> Sr 2+ withvalues estimated by Equation (26), us<strong>in</strong>g the parameters listed <strong>in</strong> Table 2. The values forArchie’s exponent n, <strong>and</strong> the sodium activity a Na were adapted to obta<strong>in</strong> a visual agreement(the Na + concentration is unknown). The Na + <strong>and</strong> Cs + data are from Kozaki et al.(1998) who removed excess salt from the (Kunipia) bentonite by dialysis after hav<strong>in</strong>g itwashed with NaCl solution, while the Sr 2+ data stem from Sato et al. (1992) who useddistilled water for saturation only <strong>and</strong> did not prepare the bentonite. This could expla<strong>in</strong>the higher activity <strong>of</strong> Na + <strong>in</strong> the experiments <strong>of</strong> Sato, but it may be that also other cationsare present <strong>in</strong> the porewater. The values for Archie’s n are low (mean<strong>in</strong>g that thetortuosity is small).


15Dashed l<strong>in</strong>es show the calculated apparent diffusion coefficient if, for Cs + the higher a Na<strong>of</strong> Sr 2+ is used, <strong>and</strong> for Sr 2+ the lower a Na <strong>of</strong> Cs + . To match the experimental data aga<strong>in</strong>,Archie’s n would have to be <strong>in</strong>creased for Cs + (to n ≈ 2), <strong>and</strong> still more reduced forSr 2+ (to n ≈ -2). The very low values for Sr 2+ suggest that Ca 2+ is present <strong>in</strong> porewater,which competes more strongly than Na + for exchange sites. For example, if only 1mmol Ca 2+ /L were added to 100 mmol Na + /L, the distribution coefficient <strong>of</strong> Sr 2+ woulddecrease 10-fold. Anyhow, it is clear that the background concentration <strong>of</strong> NaCl (whichdeterm<strong>in</strong>es a Na ) has a spectacular effect on the apparent diffusion coefficient. Unfortunately,few experiments have it well def<strong>in</strong>ed, <strong>and</strong> <strong>in</strong> most, it is not mentioned at all, norare the concentrations <strong>of</strong> other major cations known.Figure 6. Apparent diffusion coefficients measured <strong>in</strong> bentonite by Kozaki et al. (1998)<strong>and</strong> Sato et al. (1992), compared with estimates us<strong>in</strong>g Equation (26). Full l<strong>in</strong>es wereobta<strong>in</strong>ed with parameters listed <strong>in</strong> Table 2, dashed l<strong>in</strong>es for Cs + <strong>and</strong> Sr 2+ were calculatedwith the values <strong>of</strong> the background concentration <strong>of</strong> NaCl <strong>in</strong>terchanged.Table 2. Parameters for estimat<strong>in</strong>g apparent diffusion coefficients for Na + , Cs + <strong>and</strong>Sr 2+ <strong>in</strong> bentonite with Equation (26). The CEC <strong>of</strong> the bentonite was fixed to 1 eq/kg.Na Cs Srz i 1 1 2D w / (10 -9 m 2 /s) (a) 1.33 2.07 0.794(b)K I \ Na 1 15 2.8Archie's n 0.5 0.5 0.1a Na 0.1 0.1 0.3(a) Appelo <strong>and</strong> Wers<strong>in</strong> 2007(b) Appelo <strong>and</strong> Postma 2005


16In Equation (26) it is assumed that all the porosities <strong>in</strong> bentonite contribute equally todiffusion. In reality, the processes are more complicated. Water <strong>and</strong> solutes are mobile<strong>in</strong> <strong>in</strong>terlayer water, DDL water <strong>and</strong> free porewater, but diffusion coefficients are different<strong>in</strong> these doma<strong>in</strong>s. Kemper et al. (1964), Van Schaik et al. (1966) <strong>and</strong> Dufey <strong>and</strong>Laudelout (1975, 1976) measured smaller diffusion rates <strong>of</strong> HDO <strong>and</strong> cations <strong>in</strong> theDDL close to the surface <strong>of</strong> montmorillonite than farther out. Kozaki et al. (1998b;2005; 2008) measured the temperature dependence <strong>of</strong> the diffusion coefficients <strong>in</strong>bentonite <strong>and</strong> found that the activation energy varied for cations, anions <strong>and</strong> HTO withdry density: for ρ dry < 900 kg/m 3 the activation energy is similar to the one <strong>in</strong> free water,up to 1100 kg/m 3 the activation energy for Na + <strong>and</strong> Cl - is smaller than <strong>in</strong> free water, <strong>and</strong>at higher dry densities it <strong>in</strong>creases above the free water value.Generally <strong>in</strong> chemistry, if the activation energy <strong>of</strong> a process is higher than for diffusion<strong>in</strong> free water, it <strong>in</strong>dicates that chemical bonds are broken. Thus, higher activation energiesat higher bentonite densities suggest that <strong>in</strong>terlayer water, <strong>in</strong> which the cations arebonded to the bentonite surface <strong>and</strong> where water is present as hydration water ratherthan free water, is ga<strong>in</strong><strong>in</strong>g importance over diffusion <strong>in</strong> free porewater. Kozaki et al.measured mostly <strong>in</strong> distilled water <strong>in</strong> which the DDL can occupy a large portion <strong>of</strong> thepore space, caus<strong>in</strong>g the contribution <strong>of</strong> DDL water to manifest itself at <strong>in</strong>termediate drydensities.Molecular dynamics (MD) calculations <strong>in</strong>dicate that water <strong>and</strong> cations <strong>in</strong> DDL water areonly slightly less mobile than <strong>in</strong> free water (Lyklema et al. 1998; Rotenberg et al. 2007).For <strong>in</strong>terlayer water, results are dependent on the number <strong>of</strong> water layers <strong>in</strong> the structure.Sodium <strong>in</strong> the 3-layer hydrate is calculated to have diffusion coefficients that aresimilar to DDL water (Chang et al. 1995; Rotenberg et al. 2007), but <strong>in</strong> the 2- <strong>and</strong> 1-layer hydrate the diffusion coefficient <strong>of</strong> Na + seems to decrease by 1 <strong>and</strong> 2 orders <strong>of</strong>magnitude, respectively (Chang et al. 1995; Marry <strong>and</strong> Turq 1992).3.1 Models for diffusion <strong>in</strong> bentoniteCations have, relative to deuterated or tritiated water molecules, higher porewater diffusioncoefficients <strong>in</strong> bentonite <strong>and</strong> clay than <strong>in</strong> ord<strong>in</strong>ary water (Sato et al., 1992; Wers<strong>in</strong>et al. 2004b; Melkior et al. 2005; Eriksen et al. 1999). In other words, the tortuosity factorsare smaller for cations than for water molecules, <strong>and</strong> also, they differ for variouscations (Figure 6 <strong>and</strong> Table 2; Appelo <strong>and</strong> Wers<strong>in</strong> 2007). On the other h<strong>and</strong>, anionshave larger tortuosity factors (Shackelford <strong>and</strong> Daniel, 1991; Oscarson et al. 1992; Satoet al. 1992; Iversen <strong>and</strong> Jørgensen 1993; Muur<strong>in</strong>en 1994; Molera et al. 2002), which isreadily expla<strong>in</strong>ed by overlapp<strong>in</strong>g double layers at pore constrictions (Appelo <strong>and</strong>Wers<strong>in</strong>, 2007; Figure 5).Several models have been proposed for expla<strong>in</strong><strong>in</strong>g <strong>and</strong> calculat<strong>in</strong>g the different behavior<strong>of</strong> cations <strong>and</strong> anions. Ochs et al. (1998, 2001) considered the <strong>porosity</strong> to consist only<strong>of</strong> free porewater <strong>and</strong> DDL water. They calculated the enrichment <strong>of</strong> cations <strong>and</strong> thedepletion <strong>of</strong> anions <strong>in</strong> DDL water to obta<strong>in</strong> multipliers that <strong>in</strong>crease the diffusion coefficientfor cations <strong>and</strong> reduce it for anions. However, as noted earlier, diffusion <strong>in</strong> theDDL is slightly smaller than <strong>in</strong> free porewater. To compensate the <strong>in</strong>crease for cations,Ochs et al. <strong>in</strong>troduced high tortuosity factors, which rise to unexpla<strong>in</strong>ed high values


17with <strong>in</strong>creas<strong>in</strong>g density <strong>of</strong> the bentonite. As noted by Bourg et al. (2003), the escalation<strong>of</strong> the tortuosity factor is due to the neglect <strong>of</strong> <strong>in</strong>terlayer water which becomes the major<strong>porosity</strong> type when density <strong>in</strong>creases above 1800 kg/m 3 .Other models subdivide the effective diffusion coefficient <strong>in</strong> its components associatedwith free porewater, DDL water (usually denoted as surface diffusion), <strong>and</strong> <strong>in</strong>terlayerwater (Eriksen et al. 1999; Bourg et al. 2006, 2007; Kozaki et al. 2008). The effectivediffusion coefficient becomes:D e, i = ε free D p, i + ε DDL D DDL, i + ε IL D IL, i (27)which for anions turns <strong>in</strong>to (discussed <strong>in</strong> Section 5.1): freeD e ,i free D2 w,i(28)free<strong>and</strong> for cations <strong>and</strong> neutral species:freeDDL ILD e ,i free DDL ILD 222 freeDDLILw,i(29)In these models, the concentration gradient <strong>in</strong> free porewater is used as the driv<strong>in</strong>g forcefor the diffusive flux. However, the concentration scales <strong>in</strong> the diffusion doma<strong>in</strong>s aredifferent, <strong>and</strong>, <strong>of</strong> course, the concentration gradient should be calculated <strong>in</strong> accordancewith that scale. Thus, as demonstrated <strong>in</strong> section 4.2, a fractional scale must be used forcalculat<strong>in</strong>g the concentration gradients for diffusion <strong>in</strong> <strong>in</strong>terlayer water.Van Loon et al. (2007) measured the diffusion <strong>of</strong> 36 Cl - <strong>in</strong> bentonite at various NaCl concentration<strong>and</strong> obta<strong>in</strong>ed anion-accessible porosities from the data, which are assumedhere to be equal to ε free . For a comprehensive dataset, they showed that the geometricalfactor for Cl - followed Archie’s law:1G 2 nfree(30)with n = 0.9 (Figure 7). Muur<strong>in</strong>en (1994) found a similar relation, but with n = 0.6. Thel<strong>in</strong>e given by Archie’s law <strong>in</strong> Figure 7 was forced to pass through the free water value,D w , Cl = 2.03×10 -9 m 2 /s. It can be seen that this forc<strong>in</strong>g leads to a slight overestimate <strong>of</strong>the effective diffusion coefficient compared with experimental values <strong>and</strong> it may be thatthe real tortuosity is somewhat larger.


18Figure 7. The effective diffusion coefficient <strong>of</strong> Cl - <strong>in</strong> bentonite as a function <strong>of</strong> the accessible<strong>porosity</strong> (Van Loon et al. 2007). The l<strong>in</strong>e denoted as Archie’s law gives D e,Cl =(ε free ) 1.9 2.03×10 -9 m 2 /s.Also for Cl - , Leroy et al. (2006) developed a model <strong>in</strong> which the concentration <strong>in</strong> DDLwater is calculated by Donnan equilibrium:2c ,DDL , Cl 0.25 c free Cl 0.5(31)where σ is the charge <strong>of</strong> the surface to which the DDL is attached (mol/L). They nextgive the expression for the diffusion coefficient <strong>in</strong> DDL water which simplifies to:cDDL,ClDDDL, Cl Dw,Cl(32)cfree,ClAnd similar for Na + :cDDL,NaDDDL, Na Dw,Nac(33)free,NaIn DDL water, c Cl is smaller than <strong>in</strong> free water, <strong>and</strong> conversely, c Na is higher than <strong>in</strong> freewater. Thus, the model expla<strong>in</strong>s the observed trends <strong>of</strong> larger <strong>and</strong> smaller diffusion coefficientfor cations <strong>and</strong> anions, respectively. However, Leroy et al. propose Equations(32) <strong>and</strong> (33) for diffusion <strong>of</strong> a salt (NaCl) <strong>in</strong> which the denom<strong>in</strong>ator c-term stems fromthe conversion c d(ln c) = d(c), <strong>and</strong> the numerator c-term from the conversion <strong>of</strong> mobility<strong>in</strong>to diffusion coefficient. Leroy et al. assume that the concentration <strong>in</strong> the numeratorcan be split up <strong>in</strong> the constituent ions (even when the concentrations are different),while the denom<strong>in</strong>ator <strong>and</strong> the concentration gradient are taken for the salt.


19More simply, the <strong>in</strong>dividual concentrations <strong>in</strong> DDL water can be calculated by averag<strong>in</strong>gthe potential <strong>in</strong> DDL water, <strong>and</strong> calculat<strong>in</strong>g the concentration with Boltzmann’sequation: - ziFDDLcDDL, i ciexp(34) RTwhere ψ DDL is the averaged potential <strong>in</strong> DDL water. Appelo <strong>and</strong> Wers<strong>in</strong> (2007) haveshown that the averaged concentration gives an accurate approximation for the totalconcentration that follows from <strong>in</strong>tegrat<strong>in</strong>g the exponential concentration distribution <strong>in</strong>the DDL. With Equation (34), the flux through DDL water can be calculated as shown<strong>in</strong> the next section.3.2 Which concentration gradient should be used?Many authors, cited before, distribute a high proportion <strong>of</strong> the mass <strong>of</strong> cations to DDL<strong>and</strong>/or <strong>in</strong>terlayer water, <strong>and</strong> calculate the diffusive flux without realiz<strong>in</strong>g that the driv<strong>in</strong>gforce for diffusion is not the concentration, but the concentration gradient (formally, thepotential gradient). Imag<strong>in</strong>e two tubes, one filled with 1 M NaCl, the other with 0.01 MNaCl, while the concentration gradient is the same <strong>in</strong> both tubes. If the flux <strong>in</strong> the ensembleis weighed accord<strong>in</strong>g to concentration, the flux would be 99 % located <strong>in</strong> thetube with the 1 M NaCl solution. But <strong>of</strong> course, the flux is the same <strong>in</strong> the two tubes.Actually, because the activity coefficient is somewhat smaller <strong>in</strong> 1 M solution, the flux<strong>in</strong> that solution is slightly less.In free porewater <strong>and</strong> DDL water the molal scale is used for the activity <strong>of</strong> solutes(mol/kg H 2 O, divided by 1 M to arrive at a dimensionless number). The flux <strong>in</strong> freeporewater is, from comb<strong>in</strong><strong>in</strong>g Equations (20) <strong>and</strong> (29), assum<strong>in</strong>g γ i = 1:Jfree,i free dci free D2 w,i(35) dxfreeAppelo <strong>and</strong> Wers<strong>in</strong> (2007) have shown that the potential gradient <strong>in</strong> free porewater alsoapplies to diffusion <strong>in</strong> DDL water. Thus the flux through DDL water follows fromEquations (14-15), (29) <strong>and</strong> (34):JDDL,iDDLcDDL,idci DDLD2 w,i(36) c dxDDLiFor exchangeable cations <strong>in</strong> <strong>in</strong>terlayer water the molar- or equivalent-fraction is thecommonly used concentration scale. Both molar- <strong>and</strong> equivalent-fraction scales are wellestablished by experience <strong>and</strong> experiment, <strong>and</strong> expla<strong>in</strong> typical aspects <strong>of</strong> ion exchangesuch as dilution <strong>and</strong> retardation <strong>in</strong> a complex mixture <strong>of</strong> cations (Appelo <strong>and</strong> Postma,2005). The choice between the equivalent <strong>and</strong> the molar scale is mostly a matter <strong>of</strong> convenience(Ga<strong>in</strong>es <strong>and</strong> Thomas 1953). The equivalent scale has the advantage that thesum <strong>of</strong> the exchangeable cations equals the CEC (a more or less fixed number), <strong>and</strong> willbe used here.


20The equivalent fraction <strong>of</strong> an exchangeable cation is:β i = (eq i /kg) / CEC (37)Aga<strong>in</strong>, the flux <strong>in</strong> the <strong>in</strong>terlayer follows from Equations (14-15) <strong>and</strong> (29):JIL,i cIL IL,i di ILD2 w,i(38) dxILiwhere c IL,i is the concentration <strong>in</strong> the <strong>in</strong>terlayer (mol/m 3 ). For example, the Namontmorillonitediscussed before has 0.6 Na + ions <strong>in</strong> an <strong>in</strong>terlayer volume <strong>of</strong> 2.84×10 -28m 3 for the 2-layer hydrate. Divid<strong>in</strong>g by Avogadro’s number gives the concentrationc IL,Na = c IL,CEC = 3.51 kmol/m 3 . In general for exchangeable cations,c IL,i = β i (c IL,CEC ) / z i (39)<strong>and</strong> therefore: cIL,CECILdiJ IL,i ILD 2 w,iz(40)i dxIL 3.3 Example: fluxes <strong>of</strong> I - , Cs + <strong>and</strong> Sr 2+ <strong>in</strong> the 3 <strong>porosity</strong> doma<strong>in</strong>sAs an example, the concentration gradients <strong>and</strong> fluxes can be calculated <strong>in</strong> freeporewater, DDL water <strong>and</strong> <strong>in</strong>terlayer water <strong>in</strong> a column conta<strong>in</strong><strong>in</strong>g bentonite packed toa dry density <strong>of</strong> 1700 kg/m 3 . Steady state concentration differences are ma<strong>in</strong>ta<strong>in</strong>ed at thecolumn ends (0 <strong>and</strong> 1 m) <strong>of</strong> the NaCl concentration (0.3 <strong>and</strong> 0.1 M), <strong>and</strong> trace concentrations<strong>of</strong> I - , Cs + <strong>and</strong> Sr 2+ (3 <strong>and</strong> 1 μmol/L). Us<strong>in</strong>g the parameters derived fromMuur<strong>in</strong>en’s data (Table 1), the porosities are:I = 0.1 I = 0.3 Averageε IL = 0.299 - 0.0128×I 0.298 0.295 0.30ε DDL = 0.0363 / √I 0.115 0.066 0.09ε tot = 1 - 1700 / 2760 0.384 0.384 0.40 (adapted total <strong>of</strong> ε’s)ε free = ε tot - ε IL - ε DDL (-0.028) 0.023 0.01 (small positive number)For I NaCl = 0.1, the total <strong>of</strong> DDL <strong>and</strong> <strong>in</strong>terlayer water is calculated to be larger than thetotal <strong>porosity</strong>. The free <strong>porosity</strong> would be negative, which is impossible, <strong>of</strong> course, <strong>and</strong>a small positive value (0.01) is assumed <strong>in</strong> the average. In the averaged numbers, <strong>in</strong>terlayerwater constitutes 75 % <strong>of</strong> the total <strong>porosity</strong>.The total CEC = 1 eq/kg × 1.7 kg/L total volume = 1.7 eq/L. The CEC can be distributedover <strong>in</strong>terlayer water <strong>and</strong> the sum <strong>of</strong> DDL- <strong>and</strong> free porewater <strong>in</strong> proportion withspecific surface areas (Table 1). Thus, the <strong>in</strong>terlayer charge (c IL,CEC ) is (583 / (583 + 86)= 0.87) × 1.7 eq/L / (0.3 L <strong>in</strong>terlayer water/L total volume) = 4.94 eq/L <strong>in</strong>terlayer water.The DDL charge is 0.13 × 1.7 eq/L / (0.09 L DDL water/L total volume) = <strong>2.4</strong>3 eq/L.


21The tortuosity factor for the comb<strong>in</strong>ation <strong>of</strong> free porewater <strong>and</strong> DDL water is δ/ θ 2 = ε =0.1. The same factor is used for <strong>in</strong>terlayer water as well, <strong>in</strong> accordance with estimatesfrom MD calculations for 2-layer hydrated montmorillonite.The concentrations <strong>and</strong> gradients <strong>in</strong> the 3 water types can be calculated with PHREEQC<strong>and</strong> are given at x = 1 m <strong>in</strong> Table 3. The steady-state fluxes then follow us<strong>in</strong>g Equations(35, 36 <strong>and</strong> 40), for example for Cs:3152210 4 1410mol/m /s19free Cs 0.01 2.0710.(35a)10J ,J DDL , Cs219 2 43 10 0 09 2 07 10 . . . 31010. 3132210 90510mol/m /s(36a)197142IL Cs 0.3 2.071049383.2310 9.9110mol/m /s (40a)10J ,The fluxes are also listed <strong>in</strong> Table 3.The concentration gradient <strong>of</strong> Cs + <strong>in</strong> free porewater is -2×10 -3 mol/m 4 , <strong>and</strong> near-zero <strong>in</strong><strong>in</strong>terlayer water. Thus, most diffusive transport <strong>of</strong> Cs + takes place through freeporewater <strong>and</strong> DDL water. The flux <strong>in</strong> DDL water is 219 times higher than <strong>in</strong> freeporewater because the <strong>porosity</strong> is 9 times, <strong>and</strong> the concentration 24 times higher forDDL water. On the other h<strong>and</strong>, the flux <strong>of</strong> I - <strong>in</strong> DDL water is 2.7 times smaller than <strong>in</strong>free porewater because the concentration is 24 times smaller than <strong>in</strong> free porewater. Theconcentration gradient <strong>of</strong> Sr 2+ <strong>in</strong> free porewater is also -2×10 -3 mol/m 4 (negative), but <strong>in</strong><strong>in</strong>terlayer water the gradient <strong>of</strong> the exchange fraction is positive. Thus, transport <strong>of</strong> Sr 2+through <strong>in</strong>terlayer water is <strong>in</strong> opposite direction <strong>of</strong> transport through <strong>in</strong>terlayer water.The different behavior <strong>of</strong> Cs + <strong>and</strong> Sr 2+ is a consequence <strong>of</strong> the law <strong>of</strong> mass action,which favors higher charged cations on exchange sites when the free porewater solutionis diluted. The Sr 2+ concentration gradient is also positive <strong>in</strong> DDL water, but compensatedthere by the negative potential gradient which results <strong>in</strong> a positive flux. Overall,the flux through <strong>in</strong>terlayer water dom<strong>in</strong>ates, <strong>and</strong> the flux <strong>of</strong> Sr 2+ is predicted to be negative.


22Table 3. Concentrations, concentration gradients <strong>and</strong> fluxes <strong>in</strong> porewater, DDL water<strong>and</strong> <strong>in</strong>terlayer water <strong>in</strong> a column filled with bentonite packed to a dry density <strong>of</strong> 1700kg/m 3 with a steady state flux <strong>of</strong> background NaCl (gradient <strong>of</strong> -200 mol/m 4 ) <strong>and</strong> <strong>of</strong>trace concentrations <strong>of</strong> I - , Cs + <strong>and</strong> Sr 2+ . Calculations with PHREEQC (Parkhurst <strong>and</strong>Appelo, 1999).0.1 M NaCl1 μM I - , Cs + , Sr 2+ gradient ε 2D w(10 -9 m 2 /s)Flux(mol/m 2 /s)Free porewater mol/m 3 mol/m 4I - 10 -3 -2×10 -3 0.01 0.1 2.03 4.06×10 -15Cs + 10 -3 -2×10 -3 0.01 0.1 2.07 4.14×10 -15Sr 2+ 10 -3 -2×10 -3 0.01 0.1 0.794 1.59×10 -15DDL water mol/m 3 mol/m 4 or V/mI - 4.12×10 -5 -1.65×10 -4 0.09 0.1 2.03 1.50×10 -15Cs + <strong>2.4</strong>3×10 -2 -2.36×10 -5 0.09 0.1 2.07 9.05×10 -13Sr 2+ 0.59 +1.18 0.09 0.1 0.794 8.43×10 -12ψ / V 0.082 -0.051Interlayer water β (-) 1/mCs + 1.99×10 -4 -3.23×10 -7 0.3 0.1 2.07 9.91×10 -14Sr 2+ 8.11×10 -4 +1.62×10 -3 0.3 0.1 0.794 -9.51×10 -11


234 DISCUSSIONFigures 3, 4 <strong>and</strong> 6 compare model calculations <strong>and</strong> experimental data <strong>of</strong> bentonite properties,<strong>and</strong> show that models can expla<strong>in</strong> the observed trends. However, the adoptedmodel parameters are questionable <strong>in</strong> some cases (Archie’s n is very small for Sr 2+ <strong>in</strong>Table 2; the <strong>in</strong>ternal planar surface area is improbably high for Volclay <strong>in</strong> Table 1). Thediscrepancy was provisionally related to the dissolution <strong>of</strong> accessory m<strong>in</strong>erals frombentonite. This is not merely a hypothesis. <strong>Bentonite</strong> is not pure montmorillonite, it conta<strong>in</strong>sappreciable amounts <strong>of</strong> calcite <strong>and</strong> pyrite, <strong>and</strong> possibly <strong>of</strong> gypsum as well. Dissolution<strong>of</strong> gypsum may give a Ca 2+ concentration <strong>of</strong> 20 mM when it dissolves till saturation.If it does, the distribution coefficient <strong>of</strong> Sr 2+ decreases 20 times <strong>in</strong> 0.1 M NaCl solutioncompared with the same solution without Ca 2+ . For Cs + , the decrease is √20 = 4.5times. (Changes <strong>in</strong> the major concentrations always have an effect that is quadratichigher on divalent cations than on monovalent cations.) But, if only traces <strong>of</strong> gypsumare present, the Ca 2+ concentration depends on the ratio <strong>of</strong> bentonite <strong>and</strong> water, <strong>and</strong> itwill <strong>in</strong>crease with dry density <strong>of</strong> the bentonite. In that case the distribution coefficients<strong>of</strong> tracers decrease with the dry density, <strong>and</strong> the effect <strong>of</strong> decreas<strong>in</strong>g <strong>porosity</strong> <strong>and</strong> <strong>in</strong>creas<strong>in</strong>gCEC on the apparent diffusion coefficient is counterbalanced.In Figure 6, the apparent diffusion coefficient (D a ) <strong>of</strong> Sr 2+ is nearly <strong>in</strong>variant with drydensity <strong>of</strong> bentonite. Thus, <strong>porosity</strong> seems to have no effect on Sr 2+ (Archie’s n is nearzero).But, it is unlikely that the <strong>porosity</strong> effect is zero for Sr 2+ (measured by Sato et al.1992), while Cs + shows the expected trend. It suggests that the Na + <strong>and</strong>/or Ca 2+ concentrationchanges with dry density, s<strong>in</strong>ce that has different effect on the distribution coefficients<strong>of</strong> Sr 2+ <strong>and</strong> Cs + , <strong>and</strong> hence on their apparent diffusion coefficients Both Kozaki<strong>and</strong> Sato used Kunipia-bentonite, but s<strong>in</strong>ce Kozaki et al. measured twice higher D a forCs + than Sato et al. it seems that the experimental conditions were different <strong>in</strong> the experiments.Unfortunately, it is difficult to decide what factor can expla<strong>in</strong> the discrepancy,s<strong>in</strong>ce the parameters that are needed for do<strong>in</strong>g the calculations are not given (notably,the Na + <strong>and</strong> Ca 2+ concentrations for calculat<strong>in</strong>g distribution coefficients are lack<strong>in</strong>g).Another difficulty is that montmorillonite <strong>in</strong> different bentonites is not the same. Theyhave different layer charges, <strong>and</strong> the charge comes differently from the octahedral or thetetrahedral layer <strong>of</strong> the crystal, or from both. This affects swell<strong>in</strong>g differently (Ferrageet al. 2005a; Laird 2006), <strong>and</strong> thus the distribution <strong>of</strong> the porosities may change withdensity <strong>in</strong> another way. The different Cl - -accessible porosities <strong>in</strong> the experiments <strong>of</strong>Muur<strong>in</strong>en <strong>and</strong> Van Loon et al. (Figures 3 <strong>and</strong> 4, respectively) may be due to these variations,<strong>in</strong> addition to the different preparatory treatment.


24Table 4. Specific volume <strong>and</strong> relative density <strong>of</strong> water <strong>in</strong> the <strong>in</strong>terlayer space <strong>of</strong>montmorillonite. The specific volume is calculated from the measured (or calculated)number <strong>of</strong> water molecules per unit cell <strong>and</strong> the volume <strong>of</strong> the <strong>in</strong>terlayer space(a×b×(d 001 – c * )), where c * is the thickness <strong>of</strong> the TOT layer (10 -9 nm for divalent cations<strong>and</strong> K + , 0.96×10 -9 nm for lighter alkalis). The density is relative to pure water witha specific volume <strong>of</strong> 3.00×10 -29 m 3 /mol H 2 O.Authormol H 2 O/unit celld 001(nm)V sp(10 -29 m 3 /mol H 2 O)relativedensityCation(no <strong>of</strong> waterlayers)MethodPitteloud, 2000 3.9 1.20 2.90 1.03 Na (1W)NeutrondiffractionFerrage, 20070.5 1.17 0.16 0.19 Ca (1W) XRD3 1.26 4.10 0.73 Ca (1W)5.5 1.50 4.26 0.70 Ca (2W)7 1.52 3.53 0.85 Ca (2W)2.8 1.23 4.56 0.66 Na (1W) XRDFerrage,4.5 1.26 3.10 0.96 Na (1W)2005b6.4 1.54 4.29 0.70 Na (2W) XRD7 1.55 3.95 0.76 Na (2W)Karaborni,19964.30 1.25 3.19 0.94 Na (1W) Molecular7.60 1.55 3.67 0.81 Na (2W)dynamicscalculations9.10 1.83 4.52 0.66 Na (3W)However, even if the <strong>in</strong>terlayer spac<strong>in</strong>g would be known, the density <strong>of</strong> the water thereis not. Changes <strong>in</strong> water content <strong>and</strong> thickness <strong>of</strong> the <strong>in</strong>terlayer space have been measuredby weigh<strong>in</strong>g <strong>in</strong> comb<strong>in</strong>ation with XRD <strong>and</strong> neutron scatter<strong>in</strong>g, <strong>and</strong> can be calculated,<strong>in</strong> pr<strong>in</strong>ciple, by molecular dynamics simulations. Table 4 shows values <strong>of</strong> <strong>in</strong>terlayerthickness <strong>and</strong> water content from the literature. The XRD measurements are done byvary<strong>in</strong>g the relative humidity (RH) <strong>and</strong> show hysteresis, with smaller d 001 spac<strong>in</strong>g uponwett<strong>in</strong>g than with dry<strong>in</strong>g. Relative densities from Cases et al. (1992) <strong>and</strong> Bérend et al.(1995) (not <strong>in</strong> Table 4) average to about 0.65 for the 1-layer hydrate at a relative watervaporpressure <strong>of</strong> 0.4, <strong>and</strong> about 0.8 for the 2-layer hydrate, similar to the data <strong>of</strong>Ferrage et al. (2005b, 2007) listed <strong>in</strong> Table 4. In between 0 <strong>and</strong> ~90 % RH, four levelscorrespond with the completely dehydrated m<strong>in</strong>eral, <strong>and</strong> 1-, 2-, or 3-layer hydrates withNa + as <strong>in</strong>terlayer cation (Cases et al. 1992; Bérend et al. 1995; Ferrage 2005b, 2007). Inbetween, the spac<strong>in</strong>g <strong>and</strong> number <strong>of</strong> water molecules per water-layer varies gradually,suggest<strong>in</strong>g mixed-layer formation <strong>and</strong> a complicated distribution <strong>of</strong> the water over hydrationwater associated with the exchangeable cation <strong>and</strong> free water (Ferrage 2007).Table 4 shows that the density <strong>of</strong> water <strong>in</strong> the <strong>in</strong>terlayer space is smaller than <strong>of</strong> bulkwater, except for the value extracted from Pitteloud et al. (2000). When spac<strong>in</strong>g <strong>in</strong>creases,it approaches 1, but not <strong>in</strong> the MD calculations <strong>of</strong> Karaborni et al. (1996). The XRDresults are more consistent <strong>and</strong> show that water starts with hydrat<strong>in</strong>g the cation <strong>and</strong> thatthe structure exp<strong>and</strong>s then slightly. With <strong>in</strong>creas<strong>in</strong>g humidity, <strong>and</strong> as more water entersthe <strong>in</strong>terlayer space (displac<strong>in</strong>g the gas molecules from the laboratory atmosphere if notpumped vacuum), the relative density <strong>in</strong>creases until a threshold value is reached wherethe structure converts from the 1- to the 2-layer hydrate. With Na + as exchangeable cat-


25ion, the structure exp<strong>and</strong>s further <strong>in</strong>to the 3-layer hydrate when the RH <strong>in</strong>creases above85 % (Bérend et al. 1995).With <strong>in</strong>creas<strong>in</strong>g spac<strong>in</strong>g, the <strong>in</strong>terlayer water can be expected to ga<strong>in</strong> the density <strong>of</strong> freewater, as observed <strong>in</strong> the XRD studies. Thus, the decreas<strong>in</strong>g density calculated byKaraborni et al. <strong>and</strong> others (e.g. Marry <strong>and</strong> Turq, 2002) may be an artifact that is associatedwith approximations <strong>in</strong> the MD simulations (Young <strong>and</strong> Smith 2000).In the model for calculat<strong>in</strong>g the distribution <strong>of</strong> the porosities, it was assumed that thedensity is 1 kg/L (f dens = 1), <strong>and</strong> that may look erroneous given the results <strong>in</strong> Table 4.However, the XRD measurements are done at less than 90 % relative humidity, correspond<strong>in</strong>gwith activity <strong>of</strong> water less than 0.9. On the other h<strong>and</strong>, <strong>in</strong> free porewater <strong>in</strong>which the ionic strength is below 1, the activity <strong>of</strong> water is close to 1 <strong>and</strong> hence, the assumeddensity may be realistic.4.1 Experimental verification is necessary; consistency checks arepossibleThe example calculation <strong>in</strong> Section 4.3 <strong>of</strong> the fluxes <strong>of</strong> 3 differently charged tracersthrough the <strong>porosity</strong> doma<strong>in</strong>s h<strong>in</strong>ges on the parameters discussed before, but <strong>of</strong>fers atemplate for test<strong>in</strong>g by experiment. The proposed setup can be reproduced <strong>in</strong> a laboratoryexperiment by assembl<strong>in</strong>g 2 pieces <strong>of</strong> bentonite with the properties <strong>of</strong> the boundaryconditions <strong>and</strong> measur<strong>in</strong>g concentration pr<strong>of</strong>iles after given times. The bentonite needsto be cleaned from reactive accessory m<strong>in</strong>erals <strong>and</strong> washed. The result<strong>in</strong>g diffusion pr<strong>of</strong>ilescan be modeled with PHREEQC <strong>and</strong> used for verify<strong>in</strong>g model-assumptions <strong>and</strong> -parameters. In particular, the large flux <strong>of</strong> Sr 2+ through <strong>in</strong>terlayer water, which is partlya result <strong>of</strong> the tortuosity factor found <strong>in</strong> MD calculations, needs experimental validation.The MD calculations consider a s<strong>in</strong>gle, rectangular <strong>in</strong>terlayer unit <strong>and</strong> do not accountfor irregular pack<strong>in</strong>g <strong>of</strong> the units. Thus, the real tortuosity factor for cations <strong>in</strong> <strong>in</strong>terlayerwater is probably larger. Also, the same tortuosity factor was adopted for cations <strong>and</strong>anions <strong>in</strong> free porewater <strong>and</strong> DDL water. However, as illustrated <strong>in</strong> Figure 5, the tortuositymay be larger for anions than for cations <strong>and</strong> neutral species, <strong>and</strong> the viscosity <strong>of</strong>DDL water may be higher than <strong>of</strong> free porewater.Consistency checks are possible <strong>of</strong> the parameters used for DDL water. The relativeamounts <strong>of</strong> free porewater <strong>and</strong> DDL water, <strong>and</strong> the average potential <strong>of</strong> DDL watershould agree with the anion-accessible <strong>porosity</strong>, the potential that is used for calculat<strong>in</strong>gthe depletion or enrichment <strong>of</strong> anions or cations, the external surface area <strong>of</strong>montmorillonite <strong>in</strong> contact with DDL water, the thickness <strong>of</strong> DDL water, the averagepore-size, <strong>and</strong> the charge <strong>of</strong> montmorillonite that is counterbalanced <strong>in</strong> DDL water. Forexample, if the anion-accessible <strong>porosity</strong> is measured to be 0.034, the follow<strong>in</strong>g equationresults for I - :an, Ifree FDDL DDL 0. 034 exp(41) RT tottot


26Furthermore, if the flux <strong>of</strong> Cs + is 20 times the flux <strong>of</strong> HDO with the same concentrationgradient <strong>in</strong> free porewater, <strong>and</strong> the tortuosity is the same <strong>and</strong> <strong>in</strong>terlayer diffusion can beneglected:JJDDL, CsDDL, HDOD 20 Dw,Csw,HDOfree - F exp RT freeDDLDDLSimilar equations hold for anions as well as other cations.F<strong>in</strong>ally,DDL(42)ε free + ε DDL = ε tot - ε IL (43)Equations (41-43), <strong>and</strong> an estimate <strong>of</strong> ε IL , allow to obta<strong>in</strong> ε free , ε DDL <strong>and</strong> ψ DDL (<strong>and</strong> will,with the numbers selected, provide the porosities that were used <strong>in</strong> Section 4.3).In the example <strong>of</strong> Section 4.3, the concentration <strong>of</strong> I - <strong>in</strong> DDL water is 0.04 times theconcentration <strong>in</strong> free porewater, <strong>and</strong> DDL water conta<strong>in</strong>s 26 % <strong>of</strong> I - (with the ε DDL =0.09 <strong>and</strong> ε free = 0.01). Accord<strong>in</strong>gly, the assumption that free porewater equals the anionaccessible<strong>porosity</strong> is not correct: the free porewater <strong>porosity</strong> is smaller than the anionaccessible<strong>porosity</strong>. With the numbers from adequate measurements, Equations (41-43)will provide better estimates for the porosities.The DDL <strong>porosity</strong> should also be consistent with the external surface area <strong>of</strong> themontmorillonite <strong>and</strong> the thickness <strong>of</strong> DDL water. In the example <strong>of</strong> Section 4.3,bentonite with a specific external surface area <strong>of</strong> 86 m 2 /g was packed to 1700 kg/m 3 .Assum<strong>in</strong>g that the thickness <strong>of</strong> DDL water is 1 Debye length or 0.3 / √0.1 = 1 nm, theDDL <strong>porosity</strong> amounts to 1.7×10 -6 × 86 × 10 -9 m 3 /m 3 = 0.15. This is higher than thevalue (0.09) obta<strong>in</strong>ed with the <strong>porosity</strong> distribution model, but perhaps the thicknessshould be reduced, or alternatively, the specific surface area must be adapted.


275 CONCLUSIONS<strong>Diffusion</strong> <strong>of</strong> various tracers <strong>in</strong> bentonite has been measured <strong>of</strong>ten, but models that canexpla<strong>in</strong> the results are few, <strong>and</strong> unfortunately, <strong>in</strong>complete or afflicted by <strong>in</strong>correct assumptions.The behavior <strong>of</strong> montmorillonite, the m<strong>in</strong>eral that provides bentonite with itsseal<strong>in</strong>g properties, lies at the heart <strong>of</strong> the problem. From the dry state <strong>in</strong> a bentonite pellet,montmorillonite swells up <strong>in</strong> water. Concomitantly, the hydraulic conductivity isreduced to very low values, <strong>and</strong> diffusion becomes the ma<strong>in</strong>, if not the only transportmechanism. <strong>Diffusion</strong> can occur through free porewater, an electrically balanced solution,through diffuse double layer water that is present at the <strong>in</strong>terface with the m<strong>in</strong>eralsurface <strong>and</strong> has a surplus <strong>of</strong> cations, <strong>and</strong> through <strong>in</strong>terlayer water that forms part <strong>of</strong> them<strong>in</strong>eral structure. These three water types have very different diffusional properties, <strong>and</strong>the proportions they take <strong>of</strong> the total <strong>porosity</strong> vary with pack<strong>in</strong>g density <strong>and</strong> solutionchemistry. A structural model was derived that can calculate the proportions. Parameterswere derived from two experiments <strong>and</strong> satisfactory numbers were obta<strong>in</strong>ed forone, but not for the other. The different results may be due to a different orig<strong>in</strong> <strong>of</strong> thebentonites, <strong>and</strong> perhaps, to different preparation <strong>in</strong> the experiment.For assess<strong>in</strong>g the behavior <strong>of</strong> bentonite as a conf<strong>in</strong><strong>in</strong>g material for radioactive waste, amodel is needed that can calculate the diffusion <strong>of</strong> the radioactive elements <strong>in</strong> conjunctionwith environmental <strong>in</strong>teractions that change the properties <strong>of</strong> the bentonite. <strong>Diffusion</strong>through free porewater <strong>and</strong> DDL water can be calculated already with PHREEQC(Appelo <strong>and</strong> Wers<strong>in</strong>, 2007). Interlayer water is a significant part <strong>of</strong> total water <strong>in</strong>bentonite, <strong>and</strong> a module for calculat<strong>in</strong>g diffusion through <strong>in</strong>terlayer water can enhancePHREEQC’s applicability for bentonite. In PHREEQC (like <strong>in</strong> other models), thest<strong>and</strong>ard states (activity scales) are different for the (exchangeable) cations <strong>in</strong> <strong>in</strong>terlayerwater <strong>and</strong> the species <strong>in</strong> free water. Retardation by ion exchange is a crucial part <strong>of</strong> thebarrier properties <strong>of</strong> bentonite. It can be modeled easily if the concentrations <strong>of</strong> the majorions are known. Unfortunately, these have been rarely analyzed <strong>in</strong> diffusion experiments,a situation that needs to be improved to enhance confidence <strong>in</strong> the model<strong>in</strong>g.An experiment is proposed <strong>in</strong> which diffusion <strong>of</strong> tracers (HDO, I - , Cs + , Sr 2+ ) is measured<strong>and</strong> used for model verification <strong>and</strong> for determ<strong>in</strong><strong>in</strong>g model parameters. A bentoniteis to be cleaned from reactive accessory m<strong>in</strong>erals <strong>and</strong> washed. Next, 2 samples are preparedwith 2 different concentrations <strong>of</strong> background-NaCl <strong>and</strong> tracers. Sets <strong>of</strong> the 2samples are clamped together, sliced after given times, <strong>and</strong> analyzed. The experimentwill allow to assess parameters for <strong>in</strong>terlayer diffusion (for Sr 2+ predicted to be counterto the concentration gradient <strong>in</strong> free porewater), <strong>and</strong> the parameters for DDL water(fraction <strong>of</strong> free porewater, potential for calculat<strong>in</strong>g depletion <strong>and</strong> enrichment).


286 ACKNOWLEDGMENTSComments <strong>of</strong> Urs Mäder, Paul Wers<strong>in</strong> <strong>and</strong> Olivier Leup<strong>in</strong> were gratefully <strong>in</strong>corporated<strong>in</strong> this report.


29REFERENCESAppelo, C.A.J., 1977. Chemistry <strong>of</strong> water expelled from compact<strong>in</strong>g clay layers: a modelbased on Donnan equilibrium. Chem. Geol. 19, 91-98.Appelo, C.A.J. <strong>and</strong> Postma, D., 2005. Geochemistry, groundwater <strong>and</strong> pollution, 2 nd ednBalkema, Leiden, 649 p.Appelo, C.A.J. <strong>and</strong> Wers<strong>in</strong>, P., 2007. Multicomponent diffusion model<strong>in</strong>g <strong>in</strong> clay systemswith application to the diffusion <strong>of</strong> tritium, iodide <strong>and</strong> sodium <strong>in</strong> Opal<strong>in</strong>us Clay,Environ. Sci. Technol. 41, 5002-5007.Appelo, C.A.J., V<strong>in</strong>sot, A., Mettler, S. <strong>and</strong> Wechner, S., 2008. Obta<strong>in</strong><strong>in</strong>g the porewatercomposition <strong>of</strong> a clay rock by model<strong>in</strong>g the <strong>in</strong>- <strong>and</strong> out-diffusion <strong>of</strong> anions <strong>and</strong> cationsfrom an <strong>in</strong>-situ experiment. J. Contam. Hydrol., <strong>in</strong> press.Bérend, I., Cases, J.-M., François, M., Uriot, J.-P., Michot, L.J., Masion, A. <strong>and</strong> Thomas,F., 1995. Mechanism <strong>of</strong> adsorption <strong>and</strong> desorption <strong>of</strong> water vapor by homoionicmontmorillonites. 2: The Li+, Na+, K+, Rb+ <strong>and</strong> Cs+ exchanged forms. Clays <strong>and</strong> ClayM<strong>in</strong>erals 43, 324-336.Bolt, G.H. <strong>and</strong> De Haan, F.A.M., 1982. Anion exclusion <strong>in</strong> soil. In: Bolt, G.H., Ed.),Soil Chemistry: B. Physico-chemical Models. Elsevier, Amsterdam.Bourg, I.C., Sposito, G. <strong>and</strong> Bourg, A.C.M., 2003. Model<strong>in</strong>g diffusion <strong>and</strong> adsorption <strong>in</strong>compacted bentonite: a critical <strong>review</strong>. J. Contam. Hydrol. 61, 293-302.Bourg, I.C., Sposito, G. <strong>and</strong> Bourg, A.C.M., 2006. Tracer diffusion <strong>in</strong> compacted, water-saturatedbentonite. Clays Clay M<strong>in</strong>. 54, 363–374.Bourg, I.C., Sposito, G. <strong>and</strong> Bourg, A.C.M., 2007. Model<strong>in</strong>g cation diffusion <strong>in</strong> compactedwater-saturated sodium bentonite at low ionic strength. Environ. Sci. Technol.41, 8118–8122.Bradbury, M.H. <strong>and</strong> Baeyens, B., 2002. Porewater chemistry <strong>in</strong> compacted re-saturatedMX-80 bentonite – physico-chemical characterization <strong>and</strong> geochemical model<strong>in</strong>g. PSIBer. 02-10, Paul Scherrer Institut, Villigen, Switzerl<strong>and</strong>.Cases, J.M., Bérend, I., Besson, G., François, M., Uriot, J.P., Thomas, F. <strong>and</strong> Poirier,J.E., 1992. Mechanism <strong>of</strong> adsorption <strong>and</strong> desorption <strong>of</strong> water vapor by homoionicmontmorillonite: 1. The sodium exchanged form. Langmuir 8, 2730-2739.Cases, J.M., Bérend, I., François, M., Uriot, J.P., Michot, L.J. <strong>and</strong> Thomas, F., 1997.Mechanism <strong>of</strong> adsorption <strong>and</strong> desorption <strong>of</strong> water vapour by homoionicmontmorillonite: 3. The Mg2+, Ca2+, Sr2+ <strong>and</strong> Ba2+ exchanged forms. Clays Clay M<strong>in</strong>er., 45,8-22.Chang, F.-R., Skipper, N.T. <strong>and</strong> Sposito, G., 1995. Computer simulation <strong>of</strong> <strong>in</strong>terlayermolecular structure <strong>in</strong> sodium montmorillonite hydrates. Langmuir 11, 2734-2741.


30Doherty, J., 1994. PEST, Model-<strong>in</strong>dependent parameter estimation.http://www.sspa.com/pest/download.htmlDufey, J.E. <strong>and</strong> Laudelout, H.G., 1975. Self-diffusion <strong>of</strong> sodium on clay surface as <strong>in</strong>fluencedby two other alkali cations. J. Coll. Interf. Sci. 52, 340–344.Dufey, J.E., Ban<strong>in</strong>, A., Laudelout, H.G. <strong>and</strong> Chen, Y., 1976. Particle shape <strong>and</strong> sodiumself-diffusion coefficient <strong>in</strong> mixed sodium-calcium montmorillonite. Soil Sci. Soc. Am.Proc. 40, 310-314.Eriksen, T.E., Jansson, M. <strong>and</strong> Molera, M., 1999. Sorption effects on cation diffusion <strong>in</strong>compacted bentonite. Eng. Geol. 54, 231-236.Ferrage, E., Lanson, B., Malikova, N., Plançon, A., Sakharov, B.A., <strong>and</strong> Drits, V.A.,2005a. New <strong>in</strong>sights on the distribution <strong>of</strong> <strong>in</strong>terlayer water <strong>in</strong> bi-hydrated smectite fromX-ray diffraction pr<strong>of</strong>ile model<strong>in</strong>g <strong>of</strong> 00l reflections. Chem. Mater. 17, 3499-3512.Ferrage, E., Lanson, B., Sakharov, B.A., <strong>and</strong> Drits, V.A., 2005b. Investigation <strong>of</strong>smectite hydration properties by model<strong>in</strong>g <strong>of</strong> X-ray diffraction pr<strong>of</strong>iles. Part 1.Montmorillonite hydration properties. Am. M<strong>in</strong>eral., 90, 1358-1374.Ferrage, E., Kirk, C.A., Cressey, G., <strong>and</strong> Cuadros, J., 2007. Dehydration <strong>of</strong> Camontmorilloniteat the crystal scale. Part 1. Structure evolution. Am. M<strong>in</strong>eral., 92, 994-1006.Ga<strong>in</strong>es, G.L. <strong>and</strong> Thomas, H.C., 1953. Adsorption studies on clay m<strong>in</strong>erals. II. A formulation<strong>of</strong> the thermodynamics <strong>of</strong> exchange adsorption. J. Chem. Phys. 21, 714-718.García-Gutiérrez, M., Cormenzana, J.L., Missana, T. <strong>and</strong> M<strong>in</strong>garro, M., 2004. <strong>Diffusion</strong>coefficients <strong>and</strong> accessible <strong>porosity</strong> for HTO <strong>and</strong> 36 Cl <strong>in</strong> compacted FEBEX bentonite.Appl. Clay Sci. 26, 65-73.Grathwohl, P., 1998. <strong>Diffusion</strong> <strong>in</strong> porous media. Spr<strong>in</strong>ger, Berl<strong>in</strong>, 207 p.Iversen, N. <strong>and</strong> Jørgensen, B.B., 1993. <strong>Diffusion</strong> coefficients <strong>of</strong> sulfate <strong>and</strong> methane <strong>in</strong>mar<strong>in</strong>e sediments: Influence <strong>of</strong> <strong>porosity</strong>. Geochim. Cosmochim. Acta 57, 571-578.Karaborni, S., Smit, B., Heidug, W., Urai, J. <strong>and</strong> Van Oort, E., 1996. The swell<strong>in</strong>g <strong>of</strong>clays: molecular simulations <strong>of</strong> the hydration <strong>of</strong> montmorillonite. Science 271, 1102 –1104.Kemper, W.D., Maasl<strong>and</strong>, D.E.L. <strong>and</strong> Porter, L., 1964. Mobility <strong>of</strong> water adjacent tom<strong>in</strong>eral surfaces. Soil Sci. Soc. Am. Proc. 28, 164-167.Kim, H., Suk, T., Park, S. <strong>and</strong> Lee, C., 1993. Diffusivities for ions through compactedNa-bentonite with vary<strong>in</strong>g dry bulk density. Waste Management. 13, 303–308.Kozaki, T., Fujishima, A., Saito, N., Sato, S. <strong>and</strong> Ohashi, H., 1998a. Self-diffusion <strong>of</strong>sodium ions <strong>in</strong> compacted sodium montmorillonite. Nucl. Technol. 121, 63-68.


31Kozaki, T., Saito, N., Fujishima, A., Sato, S. <strong>and</strong> Ohashi, H., 1998b. Activation energyfor diffusion <strong>of</strong> chloride ions <strong>in</strong> compacted sodium montmorillonite. J. Contam. Hydrol.35, 67–75.Kozaki, T., Inada, K., Sato, S. <strong>and</strong> Ohashi, H., 2001. <strong>Diffusion</strong> mechanisms <strong>of</strong> chlorideions <strong>in</strong> sodium montmorilonite. J. Contam. Hydrol. 47, 159–170.Kozaki, T., Fujishima, A., Saito, N., Sato, S. <strong>and</strong> Ohashi, H., 2005. Effects <strong>of</strong> dry density<strong>and</strong> exchangeable cations on the diffusion process <strong>of</strong> sodium ions <strong>in</strong> compactedmontmorillonite. Eng. Geol. 81, 246–254.Kozaki, T., Liu, J. <strong>and</strong> Sato, S., 2008. <strong>Diffusion</strong> mechanism <strong>of</strong> sodium ions <strong>in</strong> compactedmontmorillonite under different NaCl concentration. Phys. Chem. Earth, <strong>in</strong> press.Laird, D.A., Thompson, M.L. <strong>and</strong> Scott, A.D., 1989. Technique for transmission electronmicroscopy <strong>and</strong> X-ray powder diffraction analyses <strong>of</strong> the same clay m<strong>in</strong>eral specimen.Clays Clay M<strong>in</strong>er. 37, 280–282.Laird, D.A., 2006. Influence <strong>of</strong> layer charge on swell<strong>in</strong>g <strong>of</strong> smectites. Appl. Clay Sci.34, 74-87.Leroy, P., Revil, A. <strong>and</strong> Coelho, D., 2006. <strong>Diffusion</strong> <strong>of</strong> ionic species <strong>in</strong> bentonite. J.Coll. Interf. Sci. 296, 248–255.Lyklema, J., Rovillard, S. <strong>and</strong> De Con<strong>in</strong>ck, J., 1998. Electrok<strong>in</strong>etics: the properties <strong>of</strong>the stagnant layer unraveled. Langmuir 14, 5659-5663.Madsen, F.T., 1998. Clay m<strong>in</strong>eralogical <strong>in</strong>vestigations related to nuclear waste disposal.Clay M<strong>in</strong>er. 33, 109-129.Marry V. <strong>and</strong> Turq P., 2003. Microscopic simulations <strong>of</strong> <strong>in</strong>terlayer structure <strong>and</strong> dynamics<strong>in</strong> bihydrated heteroionic montmorillonites. J. Phys. Chem. B 107, 1832–1839.Melkior, T., Yahiaouia, S., Motelliera, S., Thobya, D. <strong>and</strong> Tevissen, E., 2005. Cesiumsorption <strong>and</strong> diffusion <strong>in</strong> Bure mudrock samples. Appl. Clay Sci. 29, 172– 186.Molera, M. <strong>and</strong> Eriksen, T., 2002. <strong>Diffusion</strong> <strong>of</strong> 22 Na + , 85 Sr 2+ , 134 Cs + <strong>and</strong> 57 Co 2+ <strong>in</strong>bentonite clay compacted to different densities: experiments <strong>and</strong> model<strong>in</strong>g. Radiochim.Acta 90, 753–760.Muur<strong>in</strong>en, A., 1994. <strong>Diffusion</strong> <strong>of</strong> anions <strong>and</strong> cations <strong>in</strong> compacted sodium bentonite.Pub. 168, VTT, Espoo, F<strong>in</strong>l<strong>and</strong>.Muur<strong>in</strong>en, A., 2006. Ion concentration caused by an external solution <strong>in</strong>to the porewater<strong>of</strong> compacted bentonite. VTT, Espoo, F<strong>in</strong>l<strong>and</strong>.Muur<strong>in</strong>en, A., Karnl<strong>and</strong>, O. <strong>and</strong> Lehiko<strong>in</strong>en, J., 2004. Ion concentration caused by anexternal solution <strong>in</strong>to the porewater <strong>of</strong> compacted bentonite. Phys. Chem. Earth 29,119–127.


32Muur<strong>in</strong>en, A., Karnl<strong>and</strong>, O. <strong>and</strong> Lehiko<strong>in</strong>en, J., 2007. Effect <strong>of</strong> homogenization on themicrostructure <strong>and</strong> exclusion <strong>of</strong> chloride <strong>in</strong> compacted bentonite. Phys. Chem. Earth 32,485-490.Nagra, 2002. Project Opal<strong>in</strong>us Clay, Safety Report, Nagra NTB 02-05, Wett<strong>in</strong>gen,Switzerl<strong>and</strong>.Norrish, K., 1954. The swell<strong>in</strong>g <strong>of</strong> montmorillonite. Disc. Faraday Soc. 18, 120–134.Norrish, K. <strong>and</strong> Quirk, J. P. 1954. Crystall<strong>in</strong>e swell<strong>in</strong>g <strong>of</strong> montmorillonite--Use <strong>of</strong> electrolytesto control swell<strong>in</strong>g. Nature 173, 255-256.Ochs, M., Boonekamp, M., Wanner, H., Sato, H. <strong>and</strong> Yui, M., 1998. A quantitativemodel for ion diffusion <strong>in</strong> compacted bentonite. Radiochim. Acta 82, 437–443.Ochs, M., Lothenbach, B., Wanner, H., Sato, H. <strong>and</strong> Yui, M., 2001. An <strong>in</strong>tegrated sorption-diffusionmodel for the calculation <strong>of</strong> consistent distribution <strong>and</strong> diffusion coefficients<strong>in</strong> compacted bentonite. J. Contam. Hydrol. 47, 283-296.Oscarson, D.W., 1994. Surface diffusion: is it an important transport mechanism <strong>in</strong>compacted clays?. Clays Clay M<strong>in</strong>er. 42, 534–543.Oscarson, D.W., Hume, H.B., Sawatsky, N.G. <strong>and</strong> Cheung, S.C.H., 1992. <strong>Diffusion</strong> <strong>of</strong>iodide <strong>in</strong> compacted bentonite. Soil Sci. Soc. Am. J. 56, 1400– 1406.Parkhurst, D.L. <strong>and</strong> Appelo, C.A.J., 1999. User’s guide to PHREEQC (version 2). USGeol. Surv. Water Resour. Inv. Rep. 99-4259.Pitteloud, C., Powell, D.H., Soper, A.K. <strong>and</strong> Benmore, C.J., 2000. The structure <strong>of</strong> <strong>in</strong>terlayerwater <strong>in</strong> Wyom<strong>in</strong>g montmorillonite studied by neutron diffraction with isotopicsubstitution. Physica B 276-278, 236-237.<strong>Posiva</strong>, 2008. Safety case plan, 86 p. www.posiva.fi.Pusch, R., 2001. The microstructure <strong>of</strong> MX-80 clay with respect to its bulk physicalproperties under different environmental conditions. SKB, TR-01-08, 111 p.Rotenberg, B., Marry, V., Vuilleumier, R., Malikova, N., Simon C. <strong>and</strong> Turq, P., 2007.Water <strong>and</strong> ions <strong>in</strong> clays: Unravel<strong>in</strong>g the <strong>in</strong>terlayer/micropore exchange us<strong>in</strong>g moleculardynamics. Geochim. Cosmochim. Acta 71, 5089-5101.Sato, H., Ashida, T., Kohara, Y., Yui, M. <strong>and</strong> Sasaki, M., 1992. Effect <strong>of</strong> dry density ondiffusion <strong>of</strong> some radionuclides <strong>in</strong> compacted sodium bentonite. J. Nucl. Sci. Technol.29, 873-882.Sato, H. <strong>and</strong> Suzuki, S., 2003. Fundamental study on the effect <strong>of</strong> an orientation <strong>of</strong> clayparticles on diffusion pathway <strong>in</strong> compacted bentonite. Appl. Clay Sci. 23, 51-60.Shackelford, C.D. <strong>and</strong> Daniel, D.E., 1991. <strong>Diffusion</strong> <strong>in</strong> saturated soil. II: Results forcompacted clay. J. Geotechn. Eng. 117, 485-506.


33Slade, P.G., Quirk, J.P., Norrish, K., 1991. Crystall<strong>in</strong>e swell<strong>in</strong>g <strong>of</strong> smectite samples <strong>in</strong>concentrated NaCl solutions <strong>in</strong> relation to layer charge. Clays Clay M<strong>in</strong>er. 39, 234–238.Sposito, G., 1984. The surface chemistry <strong>of</strong> soils. Oxford UP, New York, 234 p.Van Brakel, J. <strong>and</strong> Heertjes, P.M., 1974. Analysis <strong>of</strong> diffusion <strong>in</strong> macroporous media <strong>in</strong>terms <strong>of</strong> a <strong>porosity</strong>, a tortuosity <strong>and</strong> a constrictivity factor. Int. J. Heat Mass Transfer 17,1093–1103.Van Loon, L.R., Glaus, M.A. <strong>and</strong> Müller, W., 2007. Anion exclusion effects <strong>in</strong> compactedbentonites: Towards a better underst<strong>and</strong><strong>in</strong>g <strong>of</strong> anion diffusion. Appl. Geochem.22, 2536-2552.Van Schaik, J.C., Kemper, W.D. <strong>and</strong> Olsen, S.R., 1966. Contribution <strong>of</strong> adsorbed cationsto diffusion <strong>in</strong> clay-water systems. Soil Sci. Soc. Am. Proc. 30, 17-22.Wers<strong>in</strong>, P., Curti, E. <strong>and</strong> Appelo, C.A.J., 2004. Modell<strong>in</strong>g bentonite-water <strong>in</strong>teractionsat high soli/liquid ratios: swell<strong>in</strong>g <strong>and</strong> diffuse double layer effects. Appl. Clay Sci. 26,249-257.Wers<strong>in</strong>, P., Van Loon, L.R., Soler, J.M., Yllera, A., Eikenberg, J., Gimmi, Th., Hernán,P. <strong>and</strong> Boisson, J.-Y., 2004b. Long-term diffusion experiment at Mont Terri: first resultsfrom field <strong>and</strong> laboratory data. Appl. Clay Sci. 26, 123-135.Young, D.A. <strong>and</strong> Smith, D.E., 2000. Simulations <strong>of</strong> clay m<strong>in</strong>eral swell<strong>in</strong>g <strong>and</strong> hydration:dependence upon <strong>in</strong>terlayer ion size <strong>and</strong> charge. J. Phys. Chem. B 104, 9163-9170.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!