• Aucun résultat trouvé

Modeling ion transport mechanisms in unsaturated porous media

N/A
N/A
Protected

Academic year: 2021

Partager "Modeling ion transport mechanisms in unsaturated porous media"

Copied!
25
0
0

Texte intégral

(1)

Publisher’s version / Version de l'éditeur:

Encyclopedia of Surface and Colloid Science, pp. 3065-3074, 2006-06-01

READ THESE TERMS AND CONDITIONS CAREFULLY BEFORE USING THIS WEBSITE.

https://nrc-publications.canada.ca/eng/copyright

Questions? Contact the NRC Publications Archive team at

PublicationsArchive-ArchivesPublications@nrc-cnrc.gc.ca. If you wish to email the authors directly, please see the first page of the publication for their contact information.

NRC Publications Archive

Archives des publications du CNRC

Access and use of this website and the material on it are subject to the Terms and Conditions set forth at

Modeling ion transport mechanisms in unsaturated porous media

Marchand, J.; Samson, E.; Snyder, K. A.; Beaudoin, J. J.

https://publications-cnrc.canada.ca/fra/droits

L’accès à ce site Web et l’utilisation de son contenu sont assujettis aux conditions présentées dans le site LISEZ CES CONDITIONS ATTENTIVEMENT AVANT D’UTILISER CE SITE WEB.

NRC Publications Record / Notice d'Archives des publications de CNRC: https://nrc-publications.canada.ca/eng/view/object/?id=007e695a-34d1-41dc-8216-3ce5c5453fe7 https://publications-cnrc.canada.ca/fra/voir/objet/?id=007e695a-34d1-41dc-8216-3ce5c5453fe7

(2)

http://irc.nrc-cnrc.gc.ca

M ode ling ion t ra nspor t m e cha nism s in

unsat urat e d porous m e dia

M a r c h a n d , J . ; S a m s o n , E . ;

S n y d e r , K . A . ; B e a u d o i n , J . J .

N R C C - 4 8 7 2 0

A version of this document is published in / Une version de ce

document se trouve dans: Encyclopedia of Surface and Colloid

Science, v. 4, June 2006, pp. 3065-3074

(3)

Modeling ion transport mechanisms in unsaturated

porous media

J. Marchand

12

E. Samson

12

K.A. Snyder

3

J.J. Beaudoin

4

1CRIB  Département de génie civil

Université Laval, Ste-Foy (Qc), Canada, G1K 7P4

2SIMCO Technologies inc.

1400, boul. du Parc Technologique, Québec (Qc), Canada, G1P 4R7

3 Building and Fire Research Laboratory

National Institute of Standards and Technology, Gaithersburg, MD 20899, USA

4Materials Laboratory - Institute for Research in Construction,

National Research Council, Ottawa (On), Canada, K1A 0R6

July 2, 2004

INTRODUCTION

The transport of ions in colloids, granular and consolidated porous media is important to a wide variety of environmental and engineering problems. Typical examples are the transport of contaminants in marine sediments, the containment of hazardous waste in soils, ionic exchange in biological membranes and processes related to the durability of building materials [14]. Comprehensive analyses [5, 6] and the development of new mod-eling techniques [7, 8] have largely contributed to the improvement of the knowledge of the various phenomena that control the transport of ions at the pore scale and at the

(4)

scale of the material.

This contribution is an attempt to briey summarize the research on the mechanisms of multi-ionic species transport in reactive and unsaturated porous media. The math-ematical description of the mass conservation equations in isothermal conditions at the local (nanoscopic) scale are discussed rst. Since, in many practical cases, porous solids are often (partially) saturated with concentrated electrolytes, emphasis is placed on the behavior of non-ideal ionic solutions. Techniques to average the equations over a Rep-resentative Elementary Volume (REV) of the material are then presented. The main hypotheses underlying the application of the homogenization approach are emphasized. Problems related to the treatment of chemical and physical interaction phenomena are also reviewed. Finally, the application of the formation factor concept and the relative importance of surface (grain-boundary) diusion phenomena are discussed.

IONIC TRANSPORT AT THE PORE SCALE

The transport of ionic species in a porous solid occurs in the liquid phase of the pore space in which ions are free to move. This includes both the bulk liquid, where ions are free to move, and the physi-sorbed liquid on the pore wall surface, where ion motion is partially constrained. A schematic representation of the pore space of a material is given in Figure 1. For this specic example, the surface of the material in contact with the pore solution is assumed to bear a negative charge. This is often the case for calcium silicate hydrates and clay minerals [9, 10]. This surface charge is usually due to a combination of the intrinsic charge of the mineral network and to surface reactions at the solid/liquid interface.

As shown in Figure 1, counterions are directly adsorbed on the surface of the solid. This layer of ions, called the Stern layer, only partially compensates the surface charge of the solid. The remaining portion of the surface charge is neutralized by an excess of positive charges is distributed in a diuse layer, forming the so-called double layer. The thickness

(5)

of the double layer is generally assumed to be small compared to the characteristics dimensions of the pore space [6]. Ions located in the free water are considered to be unaected by the surface charge of the solid.

The separation between the Stern layer and the diuse layer is called the shear plane. Ions located beyond that plane are considered to be mobile (i.e. relatively free to drift in the pore solution) [11]. In many practical cases, the transport of ions in a porous material will occur predominantly beyond the shear plane (i.e. in the electrical diuse layer and in the free water). Evidence of ionic transport at the solid particle/solution interface (i.e. in the Stern layer) has also been reported [5, 6]. The relative inuence of the latter phenomenon on the description of the transport mechanisms at the macroscopic scale will be discussed in more detail in the last section of this paper.

In many practical cases, the pore space of the solid is not fully saturated by the liquid phase. Parts of the pore space can be occupied by a gas, such as water vapor (see Figure 2). Mechanical instability at the liquid-vapor interface due to surface tension may trigger the transport of the liquid phase by capillary suction. This movement of the liquid may have a signicant inuence on the distribution of ions throughout the material.

Numerous authors have attempted to develop ionic transport models in porous solids on the basis of more or less detailed descriptions of the microstructure of the materials. Given the intrinsic complexity of most porous systems, the eective application of these models constitutes a formidable task [12]. As emphasized by Revil [6], these models usually oversimplify the microstructure by considering pores to be formed by capillaries or parallel plates. Consequently, many of these models rely on tting parameters (determined on the basis of simplied assumptions which may not physically represent the system well) to fully describe the intricate nature of the microporous solid.

Other authors have elected to average the variables and equations of interest over a Rep-resentative Elementary Volume (REV). The main advantage of this approach (called the

(6)

homogenization technique) is that it does not require any detailed knowledge of the mate-rial inner structure. Furthermore, the new averaged variables appearing in the equations represent the quantities measured in practice.

The homogenization process rst requires a comprehensive description at the local (pore) scale of the transport mechanisms of the multi-ionic species system. As previously men-tioned, various phenomena can contribute to the transport of ions in solution. For in-stance, ions will move under the inuence of an electrochemical potential µ and a bulk uid velocity v. The ux ji of a given ionic species i is approximated by an independent

sum of both phenomena [13]:

ji = −Bcigradµi+ civ (1)

The quantity ci is the concentration and B is a phenomenological transport coecient.

This formulation assumes that cross terms (a tensor transport coecient) are either neg-ligible or nonexistent. The electrochemical potential is dened as:

µi = µoi + RT ln(γici) + ziF ψ (2)

where µo

i is the constant standard potential, R is the ideal gas constant, T is the absolute

temperature, γi is the chemical activity coecient, ci is the concentration of the ionic

species i, zi is the valence number, F is the Faraday constant, and ψ is the electrical

potential.

Substituting equation (2) in equation (1) gives:

ji= −BRTgradci−BRT cigrad(ln γi) − BziF cigradψ + civ (3)

The rst term on the right-hand side of equation (3) corresponds to Fick's diusion law, which relates a ux to a concentration gradient through the diusion coecient D. Equation (3) can be modied by considering that:

B = D

p i

(7)

with Dp

i being the diusion coecient of the ionic species i, also called the self-diusion

coecient. The superscript p identies a variable dened at the scale of the pore. Re-placing equation (4) in equation (3) yields:

ji = −Dpigradci−Dpicigrad(ln γi) −

DpiziF

RT cigradψ + civ (5)

The rst term on the right-hand side of equation (5) has the identiable form of the Fick equation. The next term accounts for chemical activity eects, i.e. the non-ideal behavior of the solution. The third term accounts for the electrical coupling between the various ions. This coupling occurs when there are dierences among ionic species diusion coecients. As a result, to maintain electroneutrality, a diusion potential ψ arises to slow the faster ions and accelerate the slower ones. The rst three terms on the right-hand side of equation (5) are often referred to as the extended Nernst-Planck equation.

The coecient Dp

i appearing in equation (5) represents the self-diusion coecient of

the ionic species. In the third term on the right had side of the equation, Dp

i is used to

represent the electrochemical mobility. The self-diusion coecient can be related to the mobility ui, at innite dilution, through the Einstein relationship [3]:

ui =

DipziF

RT (6)

For a given ionic species, Dp

i is assumed constant in most ionic transport models. Values

of Dp

i for the most common ionic species, in the limit of innite dilution, can be found

in handbooks [14]. It has been argued, however, that the self-diusion concentration dependence should be inversely proportional to the solution viscosity [15]:

Dp∗i = Dpi µ ηw η

(7) where ηw is the viscosity of water and η is the viscosity of the solution. Even though this

ratio changes the value of Dp∗

i by less than 15 %, even for concentrated pore solutions, the

eect on Dp∗

i is sucient to make numerical predictions of bulk binary salt diusion agree

(8)

this viscosity correction, the activity coecient would overwhelm numerical estimates at higher concentrations.

For very dilute electrolytic solutions, the activity and electrical coupling terms in equation (5) are very small and can be neglected. However, as the ionic strength of the solution increases, these terms become more signicant. They can be viewed as correction factors to the constant diusion coecient Dp

i. This can be more easily viewed by applying the

following transformation to equation (5): ji = −Dpi µ 1 + d ln γi d ln ci + ziF RTci dψ dci ¶ gradci+ civ (8)

where the concentration dependence can be lumped into the transport coecient: Dpi′ = Dpi µ 1 + d ln γi d ln ci +ziF RTci dψ dci ¶ (9) Using the Einstein equation to approximate the electrochemical mobility ui by the

diu-sion coecient Dp

i is exact in the dilute limit, but loses accuracy at higher concentrations.

When using this approach, one has to know the concentration dependence of the mobil-ity. It has been shown [16] that the electrochemical mobility ui of ions in a solution

composed of an arbitrary mixture of ionic species can be approximated reasonably well by a relatively simple expression:

ui = u◦ i 1 + GiI 1/2 M (10) where the empirical coecients Gi range from 0.3 to 0.8 for the ionic species typically

found in pore solutions, and the quantity IM is the molar ionic strength of the solution:

IM = 1 2 N X i z2 ici (11)

where N is the number of ionic species. Using the approximations in equations (7) and (10), one can then determine the relative accuracy of approximating the mobility by the diusion coecient, for a particular application.

(9)

The mass conservation equation at the pore scale level is given by [12]: ∂ci

∂t +div(ji) + ri = 0 (12)

where ri is a source/sink term accounting for the chemical reactions undergone by the

species i in the aqueous phase. These reactions are called homogeneous [17] and involve only ions in the aqueous phase, like for example acid/base reactions. Heterogeneous reactions, which occur between the aqueous and the solid phase [17], can be divided into two distinct categories: dissolution/precipitation, and ionic exchange. At the pore scale, they can be handled by boundary conditions at the solid/liquid interface.

Substituting equation (5) in equation (12) one nds: ∂ci ∂t −div µ Dpigradci+ Dpicigrad(ln γi) + DipziF RT cigradψ − civ ¶ + ri = 0 (13)

This equation must be written for each ionic species present in the system.

The electrostatic potential (ψ) appearing in the previous equation can be calculated using Poisson's equation [3, 18]: ∇2ψ +F ǫ N X i=1 zici = 0 (14)

where N is the total number of ionic species and ǫ is the dielectric permittivity of the aqueous phase. Although the permittivity depends upon the solution ionic strength, it is typically approximated by it value for pure water.

The chemical activity coecients can be calculated with various models such as Debye-Hückel or Davies, depending on the level of concentration involved. A review of such models can be found in reference [19]. All these models relate γi to the concentration of

a given ionic species and the ionic strength of the electrolytic solution. Ionic transport calculations considering both electrical coupling between the species and chemical activity eects are presented in reference [20].

A transport process that is typically ignored in many transport models is osmosis. Osmosis is a movement of solvent from a region of lower solute concentration to a region of higher

(10)

solute concentration. The driving force of this phenomenon is a gradient of water (solvent) activity. The activity of water in a solution is given by [21]:

ln aH2O = −nM φ/mH2O (15)

where n is the number of ions in the dissolved salt (n = 2 for NaCl), M is the molality of the salt, φ is the osmotic coecient of the solution and m is the molality of water (≈ 55.6). Osmosis is also related to an osmotic pressure Π. It can be expressed as a function of the water activity aH2O. At a temperature of 300 oK, the osmotic pressure, in

units of N/m2, can be approximated using the following relationship:

Π ≈ −1.4 × 108

ln aH2O (16)

Osmotic pressure has been used to estimate concrete permeability [22, 23].

According to equation (15), the water activity for a 500 mmol/L NaCl solution is ap-proximately 0.96. The value is close to one, even at such a high concentration. This emphasizes that water activity gradients are weak for most cases of ionic transport in porous materials. Consequently, osmosis is neglected.

TRANSPORT MODEL AT THE MATERIAL SCALE

All the equations presented in the previous section were derived at the pore scale. In order to come up with a macroscopic description of the various transport processes, these equations have to be averaged over a Representative Elementary Volume (REV) of the material. In the following paragraphs, the basic concepts behind the homogenization technique are described. A comprehensive description of the technique can be found in reference [12].

As emphasized by Revil [6] and Bear [12], the dimensions of the REV should be much larger than the scales of the microscopic structure of the porous solid, and much smaller than the scale of the macroscopic phenomena. Accordingly, its size depends on the

(11)

in-trinsic properties of the material (e.g. porosity, solid matrix content, air-void content, etc.).

A REV, in which a fraction of the pore space is occupied by a gaseous phase, is shown in Figure 2. The total volume of the REV is given by Vo. The part of the volume occupied

by the liquid phase is designated by VL

o . The volumetric fraction of liquid is dened as:

θL =

VL o

Vo (17)

When the material is fully saturated, the volumetric fraction of liquid corresponds to the total porosity φ. The gaseous phase occupies a volume VG

o . It is a mixture of air and

water vapor. It is assumed that both air and water vapor ll the entire gaseous phase volume. The volumetric fraction of gas is dened as:

θG =

VG o

Vo (18)

In the remainder of the text, the superscripts L and G will designate the liquid and gaseous phases, respectively.

Let aγ denote the amount per unit volume of some extensive quantity A in the phase γ,

either solid, liquid or gas. The calculation of concentration or mass density can serve as examples for aγ. Two types of averages can be dened. The volumetric phase average is

given by: aγ = 1 Vo Z Voγ aγ dV (19)

The volumetric intrinsic phase average is dened as: aγγ = 1 Voγ Z Voγ aγ dV (20)

Using the denition of a volumetric fraction, the volumetric phase average and the volu-metric intrinsic phase average are related by the following expression:

(12)

On the basis of these denitions, as well as from the denitions of the average of a time derivative, a divergence, a gradient and a product [12], equations (13) and (14) are integrated over the REV to yield the macroscopic equations dened over the material scale. To ease the reading, uppercase symbols will be used to designate the volumetric intrinsic phase average of a given variable in the aqueous phase. It gives:

∂(θLCi) ∂t −div µ θLDigradCi+ θL DiziF RT CigradΨ + θLDiCigrad ln γi−CiV −θLc˜iv˜ L¶ + θLRi+ 1 Vo Z SLS o ji·nLS dS = 0 (22)

As can be seen, the averaging procedure introduces new parameters in the transport equations. They will be discussed in the next paragraphs.

The parameter Di is the diusion coecient at the macroscopic level. It is related to Dpi

by the expression:

Di = τLDpi (23)

where τLis the tortuosity of the aqueous phase, a purely geometrical factor accounting for

the complexity of the porous network. For uncharged porous materials (for which surface diusion phenomena can be neglected), the diusion paths of the various ionic species present in the system all have the same tortuosity in the interconnected pore space [12]. In this case, the tortuosity can be related to the so-called formation factor (F) by the following expression:

τL =

1

Fφ (24)

For a nonconducting porous solid saturated with an ionic solution, the formation factor is the ratio of the pore solution conductivity σp to the bulk material conductivity σb [24, 25]:

F = σp

σb (25)

For granular porous materials, the formation factor can be approximated by a power relationship called Archie's law [6, 25]:

(13)

where m is the cementation exponent. For dense porous materials near a percolation threshold, the formation factor can be calculated using the following expression [6]:

F = (φ − φp)−m (27)

where φp stands for the percolated porosity. In practice, the value of the cementation

exponent m usually varies from 1.5 to 2 [6, 25]. It should be emphasized that the previous denition of the tortuosity is only valid for saturated systems. Tortuosity values are likely to change with a reduction of the degree of saturation of the material (particularly when the degree of saturation is below 60 percent) [6]. It is also subjected to the local change in porosity if dissolution or precipitation reactions occur. A more complete discussion on the tortuosity factor can be found in references [12, 26].

The homogenization technique also generates the parameter γi in equation (22). This new

parameter corresponds to the chemical activity coecient calculated with the averaged concentrations Ci.

The average of the divergence of the ux leads to the surface integral in equation (22). The vector nLS is an outward (to the L-phase) unit vector on the solid/aqueous phase

interface (designated as SLS). The term (ji · nLS) gives the amount of ions crossing

the solid/aqueous phase interface, as a result of dissolution/precipitation or ion exchange reactions. However, this integral might prove dicult to evaluate. It is possible to express it dierently by performing the averaging operation on the ions in the solid phase [12]. The conservation equation at the microscopic scale is:

∂cis

∂t +div(jis) = 0 (28)

where the subscript s designates the solid phase. In contrast to equation (12), it is assumed that no chemical reactions occur within the solid phase, and that all precipita-tion/dissolution phenomena are taking place at the solid/aqueous phase interface. Aver-aging equation (28) over the REV leads to:

∂(θsCis) ∂t +div(θsjis s ) + 1 Vo Z SSL o jis·nSLdS = 0 (29)

(14)

where θs is the volumetric fraction of solid phase and nSL is an outward (to the S-phase)

unit vector on the solid/aqueous phase interface (designated as SSL). The integral in

equation (29) has the same value as the one in equation (22) but with an opposite sign since the ions coming out of the aqueous phase are being bound by the solid phase. Furthermore, the ux jis is zero since there is no ionic movement in this phase. This

allows one to write:

∂(θsCis) ∂t = 1 Vo Z SLS o ji·nLS dS (30)

Substituting equation (30) in equation (22) yields the following expression for the trans-port of ionic species:

∂(θsCis) ∂t + ∂(θLCi) ∂t −div µ θLDigradCi + θL DiziF RT CigradΨ + θLDiCigrad ln γi−θLCiV −θLc˜iv˜ L¶ + θLRi = 0 (31)

Equation (31) now includes two terms to model the chemical reactions. The term Ri

accounts for the homogeneous reactions, i.e. the ones involving only the aqueous phase. These chemical reactions are concerned with the formation of complexes in solution, i.e. ions made by the combination of other ions. There is also the term ∂(θsCis)/∂t to model

the exchange of ions between the aqueous and solid phases as a result of ion-exchange or dissolution/precipitation reactions. In most cases, the problem of interest occurs over months or years. By comparison, the rates of reaction are almost always shorter than the transport time scale. For those cases when the reaction is fast compared to transport, the reactions are treated like equilibrium reactions that occur instantaneously. This is called the local equilibrium assumption. Under this assumption, chemical reactions can be modeled by algebraic relationships [27]:

Km = N Y i=1 cνmi i γ νmi j with m = 1, . . . , M (Dissolution/precipitation) (32) xj = Kj−1γj−1 N Y j=1 cνji i γ νji j with j = 1, . . . , Nx (Complexation) (33)

(15)

where M is the number of solid phases, N is the number of ions, Km is the equilibrium

constant (or solubility constant) of the solid m, Kj is the equilibrium constant of the

complex j, ciis the concentration of the ionic species i, γiis its chemical activity coecient,

xj is the concentration of the complex j, νmi is the stoichiometric coecient of the ith

ionic species in the mth mineral, and νji is the stoichiometric coecient of the ith ionic

species in the jth complex.

Dierent techniques have been used to include chemical reactions in ionic transport prob-lems. In early models (see for instance Miller and Benson [28]) the equations that model the chemical reactions were solved simultaneously with the transport relationships. The current trend in reactive transport modeling is to split the transport part of the process from the chemical reactions. This method gained popularity after the publication of a paper by Yeh and Tripathi in 1989 [29] demonstrating that important CPU times could be saved by using such an uncoupled approach. Models by Grove and Wood [30], Yeh and Tripathi [31], Walter et al. [32] and Xu et al. [27] are all based on this operator splitting approach. Depending on the type of chemical reactions involved in the problem considered, dierent algorithms must be used to solve the transport equation. Discussions on this topic are found in references [33, 34].

The homogenization procedure introduces the term ˜civ˜ L

in the transport equation (22), which comes from the average of the product civ. The symbol ˜ indicates a variation

with regard to an average quantity. According to reference [12], the portion of the ux associated with this term is caused by variation of both the microscopic velocity and the concentration. It is referred to as the dispersion eect. The authors show that it can be expressed as follow:

˜ civ˜

L

= −DdispgradCi (34)

(16)

to give: ∂(θsCis) ∂t + ∂(θLCi) ∂t −div µ θLDi∗gradCi + θL DiziF RT CigradΨ + θLDiCigrad ln γi−θLCiV ¶ + θLRi = 0 (35) where D∗

i = Di+ Ddisp. This dispersion eect will be important for the cases where the

uid ow in the porous material is signicant. The vast majority of models developed to predict the movement of ions in soils neglect the individual ionic contribution of Di

and only take into account the dispersion because pressure head gradients can lead to important groundwater ows. This is the case for references [2734]. For cases where the uid ow is weak, such as for drying, the dispersion term vanishes. This is usually the case for ionic transport models in porous construction materials such as brick or concrete. The term V stands for the volumetric average of the liquid phase ow, and can be related to a Darcy ow. For cases where the capillary forces are dominant, such as in drying or capillary absorption situations, the transport of the uid phase can be modeled with Richards' equation. The development of this equation within the mathematical framework of the homogenization technique was presented by Withaker [35]. Richards' model, for variably saturated water ow in soils [27] can be written as:

³ φ∂Sw ∂h + SwSs ´∂h ∂t −div¡V¢ + w = 0 (36) V = KrKgrad(h + z) (37)

where h is the pressure head, z is the elevation, w is a uid source/sink term, φ is the porosity, K is the saturated hydraulic conductivity, Kr is the relative conductivity, Ss is

the specic elastic storage coecient, and Swis the water saturation. The latter parameter

is dened as Sw = θL/φ. The relative conductivity parameter is a function of the pressure

head h; for fully saturated media, it is equal to one. The term w can be used to account for water uptake by roots in the soil.

A similar equation is used to model water movement in porous construction materials [36] like bricks, mortars and concrete. The terms involving gravity are neglected, since they

(17)

are much less important than capillary eects in the very ne pores of these materials. The source sink/term w could be used to model water uptake by the hydration of cement, for instance. However, this eect is most of the time neglected.

∂θL

∂t −div(V) = 0 (38)

V= −DθgradθL (39)

where Dθ is moisture diusivity, accounting for water under both liquid and gaseous form.

More complex models have been developed to predict water content proles in porous media. Models by Selih et al. [37] and Coussy et al. [38] separately take into account the contribution of vapor and liquid water, as well as the interactions between both phases, to the overall water ow. However, these models include many parameters for which there is still no reliable measurement method. Accordingly, Richards' approach is still the most commonly used.

Finally, for multiionic models where the electrical coupling between the ions is considered, one must proceed to the averaging of Poisson's equation. This gives:

div (θLτLgradΨ) + θL F ǫ N X i=1 ziCi = 0 (40)

FUNDAMENTAL ASSUMPTIONS AT THE BASIS OF THE HOMOGENIZATION PROCEDURE

It should be emphasized that the validity of equations (31) and (40) rests on a series of classical assumptions. For instance, the derivation of the equation is based on the hypothesis that the porous material is in mechanical equilibrium and kept in isothermal conditions. The porous solid is an open system that can exchange ions, liquids and gases with its environment.

The validity of equations (31) and (40) also rests on a series of simplifying assumptions. It is hypothesized that the electrical charges at the surface of the solid do not have

(18)

any signicant inuence on the transport mechanisms at the local scale. As previously mentioned, it is well known that the concentration proles within a pore are disturbed as a result of the double layer eect [39]. The derivation of equation (31) is thus based on the hypothesis that these perturbations do not extend far from the interface and have only negligible eects on the transport mechanisms at a macroscopic scale. Similar assumptions are made for the diusion potential Ψ. The eect of the surface charges on its distribution across a pore is neglected.

In a very comprehensive investigation of ionic transport mechanisms in porous media, Revil [5, 6] demonstrated that the charged nature of the surface of most porous materials can have a signicant inuence on the transport of ions at the nanoscopic scale. For instance, his results indicate that surface transport phenomena in the Stern Layer tend to have a signicant inuence on the eective tortuosity of the system. Furthermore, this eect appears to be dierent for cations than for anions. Revil has proposed a very elegant model that takes into account the inuence of surface phenomena on macroscopic transport mechanisms. Unfortunately, the model is only valid for granular porous media saturated with a binary symmetric 1:1 electrolyte (such as NaCl).

Revil's study indicates that grain-boundary eects can be neglected for systems saturated with a concentrated electrolytic solutions [5, 6]. This is very fortunate because many porous systems are, in practice, saturated with relatively concentrated solutions. This conclusion is in good agreement with the conclusion of other studies [1, 2].

CONCLUSION

The mechanisms of ionic diusion in unsaturated porous media have been the subject of much attention over the past decade. Research done over this period has clearly allowed a better description of the various transport processes at the pore scale. Signicant eort has also been made towards a better description of macroscopic transport mechanisms of multi-ionic systems.

(19)

More research is needed to reliably describe the inuence of grain-boundary transport phenomena on the transport mechanisms in multi-ionic systems. Research is also needed to understand the inuence of temperature variations on the behaviors of these systems.

REFERENCES

[1] McDu, R.E.; Ellis, R.A. Determining diusion coecients in marine sediments : A laboratory study of the validity of resistivity techniques. American Journal of Science 1979, Vol. 279, 666-675.

[2] Oscarson, D.W. Surface diusion : Is it an important transport mechanisms in com-pacted clays?, Clay and Clay Minerals 1994, Vol. 42, No. 5, 534-543.

[3] Helerich, F. Ion exchange, McGraw-Hill, New York, USA, 1962.

[4] Taylor, H.F.W. Cement chemistry, Academic Press Inc., San Diego, USA, 1990. [5] Revil, A.; Glover, P.W.J. Theory of ionic-surface electrical conduction in porous

media. Physical Review B 1997, Vol. 55, No. 3, 1757-1773.

[6] Revil, A. Ionic diusivity, electrical conductivity, membrane and thermoelectric po-tentials in colloids and granular porous media : A unied model. Journal of Colloid and Interface Science 1999, Vol. 212, 503-522.

[7] Schwartz, L.M.; Martys, N.; Bentz, D.P.; Garboczi, E.J.; Torquato, S. Cross-property relations and permeability estimation in model porous media. Physical Review E 1993, Vol. 48, No. 6, 4584-4591.

[8] Martys, N.S. Diusion in partially saturated porous materials. Materials and Struc-tures 1999, Vol. 32, 555-562.

[9] Nachbaur, L.; Nkinamubunzi, P.C.; Nonat, A.; Mutin, J.C. Electrokinetic proper-ties which control the coagulation of silicate cement suspensions during early age hydration. Journal of Colloid and Interface Science 1998, Vol. 202, No. 2, 261-268.

(20)

[10] Sposito, G. The surface chemistry of soils, Oxford University Press, USA, 1984. [11] James, A.E.; Stillman, J.D.; Williams, D.J.A. Finite element solution of the equations

governing the ow of electrolyte in charged microporous membranes. Int. J. for Num. Meth. in Fluids 1995, Vol. 20, 1162-1178.

[12] Bear, J.; Bachmat, Y. Introduction to modeling of transport phenomena in porous media, Kluwer Academic Publishers, The Netherlands, 1991.

[13] Bockris, J.O'M.; Reddy, A.K.N. Modern electrochemistry - an introduction to an interdisciplinary area, Plenum Press, USA, 1970.

[14] Mills, R.; Lobo, V.M.M. Self-Diusion in Electrolyte Solutions, Elsevier, New York, 1989.

[15] Gordon, A.R. The diusion constant of an electrolyte, and its relation to concentra-tion. J. Chem. Phys. 1937, Vol. 5, 522526.

[16] Snyder, K.A.; Feng, X.; Keen, B.D.; Mason, T.O. Estimating the electrical con-ductivity of cement paste pore solutions from OH−, K+, and Na+ concentrations.

Cement and Concrete Research 2003, Vol. 33, 793-798.

[17] Rubin, J. Transport of reacting solutes in porous media: relation between mathemat-ical nature of problem formulation and chemmathemat-ical nature of reactions. Water Resources Research 1983, Vol. 19, No.5, 1231-1252.

[18] Samson, E.; Marchand, J.; Robert, J.L.; Bournazel, J.P. Modeling the mechanisms of ion diusion transport in porous media, International Journal of Numerical Methods in Engineering 1999, Vol. 46, 2043-2060.

[19] Samson, E.; Lemaire, G.; Marchand, J.; Beaudoin, J.J. Modeling chemical activity eects in strong ionic solutions. Computational Materials Science 1999, Vol. 15, No. 3, 285-294.

(21)

[20] Samson, E.; Marchand, J. Numerical solution for the extended Nernst-Planck model, Journal of Colloid and Interface Science 1999, Vol. 215, 1-8.

[21] Robinson, R.A.; Stokes, R.H. Electrolyte Solutions, 3rd ed., Butterworths, UK, 1970. [22] Bakker R.F.M. Permeability of blended cement concretes. in Fly Ash, Silica Fume, Slag & Other Mineral By-Products in Concrete, SP 79-30, V.M. Malhotra ed., Amer-ican Concrete Institute, Detroit (USA) 1983 589-605.

[23] Bager D.H., Eect of curing on the pore structure and permeability of cement-yash mortars. Durability of Concrete Structures Under Normal Outdoor Exposure 1984 (RILEM Symposium Proceedings; RILEM Committee CPC 14, Hannover, FRG) 102115.

[24] Collins, R.E. Flow of uids through porous materials, Reinhold Publishing, USA, 1961.

[25] Wong, P.Z. The statistical physics of sedimentary rock. Physics Today 1988, Vol. 41, No. 12, 24-32.

[26] Bachmat, Y.; Bear, J. Macroscopic modelling of transport phenomena in porous media. 1: The continuum approach. Transport in Porous Media 1986, Vol. 1, 213-240.

[27] Xu T., Samper J., Ayora C., Manzano M., Custodio E. Modeling of non-isothermal multi-component reactive transport in eld scale porous media ow systems, Journal of Hydrology 1999 Vol. 214, 144-164.

[28] Miller C.W., Benson L.V. Simulation of solute transport in a chemically reactive het-erogeneous system: model development and application, Water Resources Research 1983 Vol. 19, No. 2, 381-391.

(22)

[29] Yeh G.T., Tripathi V.S. A critical evaluation of recent developments in hydrogeo-chemical transport models of reactive multihydrogeo-chemical components, Water Resources Research 1989 Vol. 25, No. 1, 93-108.

[30] Grove D.B., Wood W.W. Prediction and eld verication of subsurface-water quality during articial recharge, Lubbock, Texas, Groundwater 1979 Vol. 17, No. 3, 250-257.

[31] Yeh G.T., Tripathi V.S. A model for simulating transport of reactive multispecies components: model development and demonstration, Water Resources Research 1991 Vol. 27, No. 12, 3075-3094.

[32] Walter A.L., Frind E.O., Blowes D.W., Ptacek C.J., Molson J.W. Modeling of multi-component reactive transport in groundwater. 1. Model development and evaluation, Water Resources Research 1994 Vol. 30, No. 11, 3137-3148.

[33] Kirkner, D.J.; Reeves, H. Multicomponent mass transport with homogeneous and heterogeneous chemical reactions: eect of the chemistry on the choice of numerical algorithm - 1. Theory. Water Resources Research 1988, Vol. 24, No. 10, 1719-1729. [34] Lichtner, P.C. Continuum model for simultaneous chemical reactions and mass

trans-port in hydrothermal systems. Geochimica & Cosmochimica Acta 1985, Vol. 49, 779-800.

[35] Whitaker S. Simultaneous heat, mass, and momentum transfer in porous media: a theory of drying, Advances in Heat Transfer 1977, Vol. 13, 119-203.

[36] Daian, J.-F. Condensation and isothermal water transfer in cement mortar: 1.Pore size distribution, equilibrium water condensation and imbibition. Transport in porous media 1988, Vol. 3, 563-589.

[37] Selih J., Sousa A.C.M., Bremner T.W. Moisture transport in initially fully saturated concrete during drying, Transport in porous media 1996 Vol. 24, 81-106.

(23)

[38] Mainguy M., Coussy O., Baroghel-Bouny V., Role of air pressure in drying of weakly permeable materials, Journal of Engineering Mechanics 2001, Vol. 127, No. 6, 582-592.

[39] Adamson, A.W. Physical chemistry of surfaces, 5th ed., John Wiley & Sons, USA, 1990.

(24)

Toward the center of the pore

Surface of the material

Stern layer Diffuse layer Free water

Solid phase Aqueous phase

Shear plane

Figure 1: Distribution of ions in a typical pore. The ions adsobed in the Stern layer are considered to be part of the solid phase. The ions in the free water as well as in the diuse layer are free to move and participate in the diusion process.

(25)

Solid Solid Solid Solid Liquid Gas Vo Vo Vo L G

Figure

Figure 1: Distribution of ions in a typical pore. The ions adsobed in the Stern layer are considered to be part of the solid phase
Figure 2: The Representative Elementary Volume (REV)

Références

Documents relatifs

This study has demonstrated that loading of pl-NP, specifically of amino-functional carbon dots, from dispersion state into the virtually fully interconnected void system of

The documents may come from teaching and research institutions in France or abroad, or from public or private research centers.. L’archive ouverte pluridisciplinaire HAL, est

We present a passive model learning method called COnfECt to infer such models from execution traces in which no information is provided to identify components.. We describe the

Mitra Ahmadi, Lucie Sancey, Arnaud Briat, Laurent Riou, Didier Boturyn, Pascal Dumy, Daniel Fagret, Catherine Ghezzi, Jean-Philippe Vuillez. To cite

The work was focused on the crack developing behavior of 2300 mm reinforced concrete beams, through carbon fiber reinforced polymers (CFRP) sheets..

(a) A comparison between C 10 of the papers first cited by the discoverers (red), high h-index authors (blue), and all researchers (black), each dot represents the mean C 10 of

/ La version de cette publication peut être l’une des suivantes : la version prépublication de l’auteur, la version acceptée du manuscrit ou la version de l’éditeur. Access

Moreover, the degree of realism can be tuned depending on the question to be addressed, from testing a specific hypothesis to building a full-grown artificial evolutionary