30.12.2012 Views

r - Les thèses en ligne de l'INP - Institut National Polytechnique de ...

r - Les thèses en ligne de l'INP - Institut National Polytechnique de ...

r - Les thèses en ligne de l'INP - Institut National Polytechnique de ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

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

�� ��� �� ����������� ��<br />

�����������<br />

������������������������<br />

�������������������������<br />

������������<br />

������� ��� �<br />

<strong>Institut</strong> <strong>National</strong> <strong>Polytechnique</strong> <strong>de</strong> Toulouse (INP Toulouse)<br />

� ���������� �� ���������� �<br />

Sci<strong>en</strong>ces <strong>de</strong> la terre et <strong>de</strong>s planètes soli<strong>de</strong>s (STP) transport <strong>en</strong> milieux poreux<br />

��������� �� �������� ��� �<br />

Hossein DAVARZANI<br />

�� �<br />

v<strong>en</strong>dredi 15 janvier 2010<br />

����� �<br />

Déterminations Théorique et Expérim<strong>en</strong>tale <strong>de</strong>s<br />

Coeffici<strong>en</strong>ts <strong>de</strong> Diffusion et <strong>de</strong> Thermodiffusion<br />

<strong>en</strong> Milieu Poreux<br />

����<br />

Ab<strong>de</strong>lka<strong>de</strong>r MOJTABI, Professeur à l’Université Paul Sabatier (UPS), Présid<strong>en</strong>t du jury<br />

Michel QUINTARD, Directeur <strong>de</strong> Recherche au CNRS, IMFT Directeur <strong>de</strong> thèse<br />

Manuel MARCOUX, MCF à l’Université <strong>de</strong> Picardie Jules Verne, Membre<br />

Pierre COSTESEQUE, MCF à l’Université Paul Sabatier, (UPS) Membre<br />

Christelle LATRILLE, Ingénieur <strong>de</strong> Recherche au CEA <strong>de</strong> Paris, Membre<br />

����� ��������� �<br />

Sci<strong>en</strong>ces <strong>de</strong> l'Univers, <strong>de</strong> l'Environnem<strong>en</strong>t et <strong>de</strong> l'Espace (SDU2E)<br />

����� �� ��������� �<br />

<strong>Institut</strong> <strong>de</strong> Mécanique <strong>de</strong>s Flui<strong>de</strong>s <strong>de</strong> Toulouse (IMFT)<br />

������������ �� ����� �<br />

Michel QUINTARD<br />

����������� �<br />

Azita AHMADI-SENICHAULT, Professeur à l’ENSAM <strong>de</strong> Bor<strong>de</strong>aux<br />

Ziad SAGHIR, Professeur à l’Université <strong>de</strong> Ryerson Canada


Theoretical and Experim<strong>en</strong>tal Determination of Effective<br />

Diffusion and Thermal diffusion Coeffici<strong>en</strong>ts in Porous Media<br />

Abstract<br />

A multicompon<strong>en</strong>t system, un<strong>de</strong>r nonisothermal condition, shows mass transfer with cross<br />

effects <strong>de</strong>scribed by the thermodynamics of irreversible processes. The flow dynamics and<br />

convective patterns in mixtures are more complex than those of one-compon<strong>en</strong>t fluids due<br />

to interplay betwe<strong>en</strong> advection and mixing, solute diffusion, and thermal diffusion (or<br />

Soret effect). This can modify species conc<strong>en</strong>trations of fluids crossing through a porous<br />

medium and leads to local accumulations. There are many important processes in nature<br />

and industry where thermal diffusion plays a crucial role. Thermal diffusion has various<br />

technical applications, such as isotope separation in liquid and gaseous mixtures,<br />

id<strong>en</strong>tification and separation of cru<strong>de</strong> oil compon<strong>en</strong>ts, coating of metallic parts, etc. In<br />

porous media, the direct resolution of the convection-diffusion equations are practically<br />

impossible due to the complexity of the geometry; therefore the equations <strong>de</strong>scribing<br />

average conc<strong>en</strong>trations, temperatures and velocities must be <strong>de</strong>veloped. They might be<br />

obtained using an up-scaling method, in which the complicated local situation (transport of<br />

<strong>en</strong>ergy by convection and diffusion at pore scale) is <strong>de</strong>scribed at the macroscopic scale. At<br />

this level, heat and mass transfers can be characterized by effective t<strong>en</strong>sors. The aim of this<br />

thesis is to study and un<strong>de</strong>rstand the influ<strong>en</strong>ce that can have a temperature gradi<strong>en</strong>t on the<br />

flow of a mixture. The main objective is to <strong>de</strong>termine the effective coeffici<strong>en</strong>ts mo<strong>de</strong>lling<br />

the heat and mass transfer in porous media, in particular the effective coeffici<strong>en</strong>t of thermal<br />

diffusion. To achieve this objective, we have used the volume averaging method to obtain<br />

the mo<strong>de</strong>lling equations that <strong>de</strong>scribes diffusion and thermal diffusion processes in a<br />

homog<strong>en</strong>eous porous medium. These results allow characterising the modifications<br />

induced by the thermal diffusion on mass transfer and the influ<strong>en</strong>ce of the porous matrix<br />

properties on the thermal diffusion process. The obtained results show that the values of<br />

these coeffici<strong>en</strong>ts in porous media are completely differ<strong>en</strong>t from the one of the fluid<br />

mixture, and should be measured in realistic conditions, or evaluated with the theoretical<br />

technique <strong>de</strong>veloped in this study. Particularly, for low Péclet number (diffusive regime)<br />

the ratios of effective diffusion and thermal diffusion to their molecular coeffici<strong>en</strong>ts are<br />

almost constant and equal to the inverse of the tortuosity coeffici<strong>en</strong>t of the porous matrix,<br />

II


while the effective thermal conductivity is varying by changing the solid conductivity. In<br />

the opposite, for high Péclet numbers (convective regime), the above m<strong>en</strong>tioned ratios<br />

increase following a power law tr<strong>en</strong>d, and the effective thermal diffusion coeffici<strong>en</strong>t<br />

<strong>de</strong>creases. In this case, changing the solid thermal conductivity also changes the value of<br />

the effective thermal diffusion and thermal conductivity coeffici<strong>en</strong>ts. Theoretical results<br />

showed also that, for pure diffusion, ev<strong>en</strong> if the effective thermal conductivity <strong>de</strong>p<strong>en</strong>ds on<br />

the particle-particle contact, the effective thermal diffusion coeffici<strong>en</strong>t is always constant<br />

and in<strong>de</strong>p<strong>en</strong>d<strong>en</strong>t of the connectivity of the solid phase. In or<strong>de</strong>r to validate the theory<br />

<strong>de</strong>veloped by the up-scaling technique, we have compared the results obtained from the<br />

homog<strong>en</strong>ised mo<strong>de</strong>l with a direct numerical simulation at the microscopic scale. These two<br />

problems have be<strong>en</strong> solved using COMSOL Multiphysics, a commercial finite elem<strong>en</strong>ts<br />

co<strong>de</strong>. The results of comparison for differ<strong>en</strong>t parameters show an excell<strong>en</strong>t agreem<strong>en</strong>t<br />

betwe<strong>en</strong> theoretical and numerical mo<strong>de</strong>ls. In all cases, the structure of the porous medium<br />

and the dynamics of the fluid have to be tak<strong>en</strong> into account for the characterization of the<br />

mass transfer due to thermal diffusion. This is of great importance in the conc<strong>en</strong>tration<br />

evaluation in the porous medium, like in oil reservoirs, problems of pollution storages and<br />

soil pollution transport. Th<strong>en</strong> to consolidate these theoretical results, new experim<strong>en</strong>tal<br />

results have be<strong>en</strong> obtained with a two-bulb apparatus are pres<strong>en</strong>ted. The diffusion and<br />

thermal diffusion of a helium-nitrog<strong>en</strong> and helium-carbon dioxi<strong>de</strong> systems through<br />

cylindrical samples filled with spheres of differ<strong>en</strong>t diameters and thermal properties have<br />

be<strong>en</strong> measured at the atmospheric pressure. The porosity of each medium has be<strong>en</strong><br />

<strong>de</strong>termined by construction of a 3D image of the sample ma<strong>de</strong> with an X-ray tomograph<br />

<strong>de</strong>vice. Conc<strong>en</strong>trations are <strong>de</strong>termined by a continuous analysing the gas mixture<br />

composition in the bulbs with a katharometer <strong>de</strong>vice. A transi<strong>en</strong>t-state method for coupled<br />

evaluation of thermal diffusion and Fick coeffici<strong>en</strong>ts in two bulbs system has be<strong>en</strong><br />

proposed. The <strong>de</strong>termination of diffusion and thermal diffusion coeffici<strong>en</strong>ts is done by<br />

comparing the temporal experim<strong>en</strong>tal results with an analytical solution mo<strong>de</strong>lling the<br />

mass transfer betwe<strong>en</strong> two bulbs. The results are in good agreem<strong>en</strong>t with theoretical results<br />

and emphasize the porosity of the medium influ<strong>en</strong>ce on both diffusion and thermal<br />

diffusion process. The results also showed that the effective thermal diffusion coeffici<strong>en</strong>ts<br />

are in<strong>de</strong>p<strong>en</strong>d<strong>en</strong>t from thermal conductivity ratio and particle-particle touching.<br />

III


Déterminations Théorique et Expérim<strong>en</strong>tale <strong>de</strong>s Coeffici<strong>en</strong>ts <strong>de</strong><br />

Diffusion et <strong>de</strong> Thermodiffusion Effectifs <strong>en</strong> Milieu Poreux<br />

Résumé <strong>en</strong> français<br />

<strong>Les</strong> conséqu<strong>en</strong>ces liées à la prés<strong>en</strong>ce <strong>de</strong> gradi<strong>en</strong>ts thermiques sur le transfert <strong>de</strong> matière <strong>en</strong><br />

milieu poreux sont <strong>en</strong>core aujourd’hui mal appréh<strong>en</strong>dées, ess<strong>en</strong>tiellem<strong>en</strong>t <strong>en</strong> raison <strong>de</strong> la<br />

complexité induite par la prés<strong>en</strong>ce <strong>de</strong> phénomènes couplés (thermodiffusion ou effet<br />

Soret).<br />

Le but <strong>de</strong> cette thèse est d’étudier et <strong>de</strong> compr<strong>en</strong>dre l’influ<strong>en</strong>ce que peut avoir un gradi<strong>en</strong>t<br />

thermique sur l’écoulem<strong>en</strong>t d’un mélange. L’objectif principal est <strong>de</strong> déterminer les<br />

coeffici<strong>en</strong>ts effectifs modélisant les transferts <strong>de</strong> chaleur et <strong>de</strong> matière <strong>en</strong> milieux poreux,<br />

et <strong>en</strong> particulier le coeffici<strong>en</strong>t <strong>de</strong> thermodiffusion effectif. En utilisant la technique <strong>de</strong><br />

changem<strong>en</strong>t d’échelle par prise <strong>de</strong> moy<strong>en</strong>ne volumique nous avons développé un modèle<br />

macroscopique <strong>de</strong> dispersion incluant la thermodiffusion. Nous avons étudié <strong>en</strong> particulier<br />

l'influ<strong>en</strong>ce du nombre <strong>de</strong> Péclet et <strong>de</strong> la conductivité thermique sur la thermodiffusion. <strong>Les</strong><br />

résultats ont montré que pour <strong>de</strong> faibles nombres <strong>de</strong> Péclet, le nombre <strong>de</strong> Soret effectif <strong>en</strong><br />

milieu poreux est le même que dans un milieu libre, et ne dép<strong>en</strong>d pas du ratio <strong>de</strong> la<br />

conductivité thermique (soli<strong>de</strong>/liqui<strong>de</strong>). À l'inverse, <strong>en</strong> régime convectif, le nombre <strong>de</strong><br />

Soret effectif diminue. Dans ce cas, un changem<strong>en</strong>t du ratio <strong>de</strong> conductivité changera le<br />

coeffici<strong>en</strong>t <strong>de</strong> thermodiffusion effectif. <strong>Les</strong> résultats théoriques ont montré égalem<strong>en</strong>t que,<br />

lors <strong>de</strong> la diffusion pure, même si la conductivité thermique effective dép<strong>en</strong>d <strong>de</strong> la<br />

connectivité <strong>de</strong> la phase soli<strong>de</strong>, le coeffici<strong>en</strong>t effectif <strong>de</strong> thermodiffusion est toujours<br />

constant et indép<strong>en</strong>dant <strong>de</strong> la connectivité <strong>de</strong> la phase soli<strong>de</strong>. Le modèle macroscopique<br />

obt<strong>en</strong>u par cette métho<strong>de</strong> est validé par comparaison avec <strong>de</strong>s simulations numériques<br />

directes à l'échelle <strong>de</strong>s pores. Un bon accord est observé <strong>en</strong>tre les prédictions théoriques<br />

prov<strong>en</strong>ant <strong>de</strong> l'étu<strong>de</strong> à l’échelle macroscopique et <strong>de</strong>s simulations numériques au niveau <strong>de</strong><br />

l’échelle <strong>de</strong> pores. Ceci démontre la validité du modèle théorique proposé. Pour vérifier et<br />

consoli<strong>de</strong>r ces résultats, un dispositif expérim<strong>en</strong>tal a été réalisé pour mesurer les<br />

coeffici<strong>en</strong>ts <strong>de</strong> transfert <strong>en</strong> milieu libre et <strong>en</strong> milieu poreux. Dans cette partie, les nouveaux<br />

résultats expérim<strong>en</strong>taux sont obt<strong>en</strong>us avec un système du type « Two-Bulb apparatus ». La<br />

diffusion et la thermodiffusion <strong>de</strong>s systèmes binaire hélium-azote et hélium-dioxy<strong>de</strong> <strong>de</strong><br />

carbone, à travers <strong>de</strong>s échantillons cylindriques remplis <strong>de</strong> billes <strong>de</strong> différ<strong>en</strong>ts diamètres et<br />

IV


propriétés thermiques, sont ainsi mesurées à la pression atmosphérique. La porosité <strong>de</strong><br />

chaque milieu a été déterminée par la construction d'une image 3D <strong>de</strong> l'échantillon par<br />

tomographie. <strong>Les</strong> conc<strong>en</strong>trations sont déterminées par l'analyse <strong>en</strong> continu <strong>de</strong> la<br />

composition du mélange <strong>de</strong> gaz dans les ampoules à l’ai<strong>de</strong> d’un catharomètre. La<br />

détermination <strong>de</strong>s coeffici<strong>en</strong>ts <strong>de</strong> diffusion et <strong>de</strong> thermodiffusion est réalisée par<br />

confrontation <strong>de</strong>s relevés temporels <strong>de</strong>s conc<strong>en</strong>trations avec une solution analytique<br />

modélisant le transfert <strong>de</strong> matière <strong>en</strong>tre <strong>de</strong>ux ampoules.<br />

<strong>Les</strong> résultats sont <strong>en</strong> accord avec les résultats théoriques. Cela permet <strong>de</strong> conforter<br />

l’influ<strong>en</strong>ce <strong>de</strong> la porosité <strong>de</strong>s milieux poreux sur les mécanismes <strong>de</strong> diffusion et <strong>de</strong><br />

thermodiffusion. Ce travail ouvre ainsi la voie à une prise <strong>en</strong> compte <strong>de</strong> l’<strong>en</strong>semble <strong>de</strong>s<br />

mécanismes <strong>de</strong> diffusion dans l’établissem<strong>en</strong>t <strong>de</strong>s modélisations numériques du transport<br />

<strong>en</strong> milieu poreux sous conditions non isothermes.<br />

V


Out beyond i<strong>de</strong>as of wrongdoing and rightdoing,<br />

there is a field. I will meet you there.<br />

Wh<strong>en</strong> the soul lies down in that grass,<br />

the world is too full to talk about<br />

language, i<strong>de</strong>as,<br />

ev<strong>en</strong> the phrase “each other”<br />

doesn't make any s<strong>en</strong>se.<br />

Rumi<br />

No amount of experim<strong>en</strong>tation can ever prove me right;<br />

a single experim<strong>en</strong>t can prove me wrong.<br />

Ce n’est pas parce que les choses sont difficiles que nous n’osons pas,<br />

c’est parce que nous n’osons pas qu’elles sont difficiles.<br />

Sénèque<br />

VI<br />

Einstein


Remerciem<strong>en</strong>ts<br />

Ma thèse, comme bi<strong>en</strong> d’autres, a nécessité <strong>de</strong> nombreux efforts <strong>de</strong> motivation et <strong>de</strong><br />

pati<strong>en</strong>ce, et n’aurait pu aboutir sans la contribution et le souti<strong>en</strong> d’un grand nombre <strong>de</strong><br />

personnes. Comm<strong>en</strong>t pourrais-je <strong>en</strong> effet remercier <strong>en</strong> seulem<strong>en</strong>t quelques mots les g<strong>en</strong>s<br />

qui m’ont sout<strong>en</strong>u p<strong>en</strong>dant ces trois années, tant leur ai<strong>de</strong> et leur prés<strong>en</strong>ce quotidi<strong>en</strong>ne ont<br />

été précieuses à mes yeux ?<br />

Dans un premier temps, je ti<strong>en</strong>s à remercier avec beaucoup <strong>de</strong> respect et <strong>de</strong><br />

reconnaissance Michel Quintard, mon directeur <strong>de</strong> thèse et aussi le responsable du groupe<br />

GEMP qui m’a accueilli à l’IMFT <strong>de</strong> Toulouse. Je le remercie pour sa confiance <strong>en</strong> moi, ce<br />

qui m’a permis d’effectuer cette thèse et <strong>en</strong> même temps appr<strong>en</strong>dre une langue et une<br />

culture très riche, que j’apprécie beaucoup. Sa rigueur, ainsi que ses qualités humaines tout<br />

au long <strong>de</strong> ces trois années auront très largem<strong>en</strong>t contribué à m<strong>en</strong>er à bi<strong>en</strong> ce travail. Je<br />

p<strong>en</strong>se notamm<strong>en</strong>t aux nombreuses relectures <strong>de</strong> docum<strong>en</strong>ts, mais égalem<strong>en</strong>t à l’ai<strong>de</strong> très<br />

précieuse apportée lors <strong>de</strong>s difficultés r<strong>en</strong>contrées durant cette pério<strong>de</strong>. Je souhaite ici dire<br />

particulièrem<strong>en</strong>t merci à Michel Quintard et à son épouse, Brigitte, pour m’avoir donné<br />

tellem<strong>en</strong>t d’amitié, <strong>en</strong> parallèle à un travail sérieux, d’avoir passé d’agréables mom<strong>en</strong>ts, <strong>de</strong><br />

bons repas français et pour le week-<strong>en</strong>d spéléologique qui était un mom<strong>en</strong>t inoubliable.<br />

Rester dans la nature sauvage m’a permis <strong>de</strong> souffler et <strong>de</strong> me ressourcer afin <strong>de</strong> rev<strong>en</strong>ir à<br />

ma thèse avec le cerveau libéré et les idées plus claires.<br />

Ensuite, je remercie très chaleureusem<strong>en</strong>t Manuel Marcoux, pour m’avoir <strong>en</strong>cadré et<br />

guidé au quotidi<strong>en</strong> avec une gran<strong>de</strong> adresse. Je lui suis reconnaissant pour son esprit<br />

d’ouverture, son professionnalisme, sa pédagogie, sa disponibilité ainsi que ses qualités<br />

humaines. Ses yeux d’expert tant sur le plan théorique qu’expérim<strong>en</strong>tal ont apporté<br />

beaucoup à mes travaux <strong>de</strong> recherche. Merci Manuel pour les longues heures consacrées à<br />

vérifier et corriger ces nombreux articles, prés<strong>en</strong>tations, manuscrit <strong>de</strong> thèse, et pour ton<br />

ai<strong>de</strong> et tes conseils <strong>en</strong> <strong>de</strong>hors du travail. Sincèrem<strong>en</strong>t, j’avais les meilleures <strong>en</strong>cadrants qui<br />

peuv<strong>en</strong>t exister !<br />

Une partie <strong>de</strong> ma thèse a été financé par le projet ANR Fluxobat, je ti<strong>en</strong>s donc à<br />

remercier une nouvelle fois Manuel Marcoux et Michel Quintard <strong>en</strong> tant que responsable<br />

sci<strong>en</strong>tifique <strong>de</strong> ce projet à l’IMFT et responsable du groupe GEMP, ainsi que Jacques<br />

Magnau<strong>de</strong>t, le directeur du laboratoire.<br />

Je remercie l’attaché <strong>de</strong> coopération sci<strong>en</strong>tifique et technique <strong>de</strong> l’ambassa<strong>de</strong> <strong>de</strong><br />

France à Téhéran, Sixte Blanchy, pour m’avoir attribué une bourse du gouvernem<strong>en</strong>t<br />

français p<strong>en</strong>dant un an. Je ne peux pas oublier <strong>de</strong> remercier chaleureusem<strong>en</strong>t Majid<br />

Kholghi mon professeur <strong>de</strong> Master pour son ai<strong>de</strong> p<strong>en</strong>dant la pério<strong>de</strong> <strong>de</strong>s démarches<br />

administratives p<strong>en</strong>dant l’inscription ; mais, malheureusem<strong>en</strong>t, les circonstances ne nous<br />

ont pas permis <strong>de</strong> travailler <strong>en</strong>semble. Je voudrais remercier très chaleureusem<strong>en</strong>t le<br />

responsable <strong>de</strong>s relations internationales <strong>de</strong> l’ENSEEIHT, Majid Ahmadpanah, pour son<br />

VII


assistance précieuse. Je ti<strong>en</strong>s à remercier Hadi Ghorbani, mon anci<strong>en</strong> collègue <strong>de</strong><br />

l’université <strong>de</strong> Shahrood, qui m’a toujours supporté et <strong>en</strong>couragé.<br />

Je remercie Azita Ahmadi qui m’a aidé et m’a supporté dans bi<strong>en</strong> <strong>de</strong>s situations<br />

difficiles, ainsi que pour le démarrage <strong>de</strong> la thèse.<br />

J’adresse mes sincères remerciem<strong>en</strong>ts à Ziad Saghir et Azita Ahmadi, qui ont accepté<br />

<strong>de</strong> rapporter sur ce travail. Je leur suis reconnaissant pour les remarques et comm<strong>en</strong>taires<br />

éclairés qu’ils ont pu porter à la lecture <strong>de</strong> ce manuscrit.<br />

Je remercie Ka<strong>de</strong>r Mojtabi qui m’a fait l'honneur <strong>de</strong> prési<strong>de</strong>r le jury <strong>de</strong> cette thèse.<br />

J’exprime mes profonds remerciem<strong>en</strong>ts à Christelle Latrille et Piere Costesèque pour avoir<br />

accepté <strong>de</strong> juger ce travail. Je remercie tout particulièrem<strong>en</strong>t Ka<strong>de</strong>r Mojtabi et Piere<br />

Costeseque <strong>de</strong> l’IMFT pour les discussions constructives durant ma thèse sur le sujet <strong>de</strong> la<br />

thermodiffusion. Je remercie Helmut pour sa prés<strong>en</strong>ce à ma sout<strong>en</strong>ance qui m’a donné<br />

beaucoup d’énergie.<br />

Je remercie Gérald Deb<strong>en</strong>est, Rachid Ababou et Franck Plouraboué pour avoir suivi<br />

mon travail, leurs <strong>en</strong>couragem<strong>en</strong>ts et leurs conseils constructifs.<br />

J’ai aussi eu l'honneur <strong>de</strong> r<strong>en</strong>contrer Massoud Kaviany au cours d'une <strong>de</strong> ses visites à<br />

l’IMFT, je le remercie pour ses conseils généraux qui m’ont été utiles.<br />

Merci à Juliette Chastanet, anci<strong>en</strong>ne post-doc à l’IMFT, qui m’a beaucoup aidé à<br />

compr<strong>en</strong>dre la théorie du changem<strong>en</strong>t d’échelle et qui a vérifié mes calculs numériques<br />

durant ma <strong>de</strong>uxième année <strong>de</strong> thèse.<br />

Le travail rapporté dans ce manuscrit a été réalisé à l’<strong>Institut</strong> <strong>de</strong> Mécanique <strong>de</strong>s<br />

Flui<strong>de</strong>s <strong>de</strong> Toulouse, dans le Groupe d’Etu<strong>de</strong> <strong>de</strong>s Milieux Poreux. Je ti<strong>en</strong>s donc à remercier<br />

la direction <strong>de</strong> l’IMFT, et H<strong>en</strong>ri Boisson. Je remercie égalem<strong>en</strong>t tout le personnel <strong>de</strong><br />

l’IMFT et <strong>en</strong> particulier Suzy Bernard, Yannick Exposito, Doris Barrau, Muriel Sabater,<br />

Sandrine Chupin, Hervé Ayroles. Je remercie Lionel Le Fur, le technici<strong>en</strong> du groupe pour<br />

son ai<strong>de</strong> à la mise <strong>en</strong> place du dispositif expérim<strong>en</strong>tal.<br />

Merci à David Bailly mon ami et collègue du bureau 210 p<strong>en</strong>dant <strong>de</strong>ux ans et quelques<br />

mois. Quand il n’y avait personne au laboratoire, bi<strong>en</strong> tard, il y avait toujours David et ça<br />

m’a donné <strong>en</strong>vie <strong>de</strong> rester et travailler. David, je n’oublierai jamais nos discussions sur<br />

différ<strong>en</strong>ts sujets, durant les pauses. <strong>Les</strong> débats qui comm<strong>en</strong>c<strong>en</strong>t par <strong>de</strong>s sujets sci<strong>en</strong>tifiques<br />

et souv<strong>en</strong>t se termin<strong>en</strong>t par <strong>de</strong>s sujets culturels, historiques ou bi<strong>en</strong> mystérieux. Et je<br />

remercie sa « diptite » chérie, Emma Flor<strong>en</strong>s, futur docteur <strong>de</strong> l’IMFT, qui passait souv<strong>en</strong>t<br />

pour nous voir.<br />

Je remercie aussi mon amie et ma collègue <strong>de</strong> bureau, Marion Musielak, anci<strong>en</strong>ne<br />

stagiaire et nouvelle doctorante très sérieuse. Je la remercie pour ses <strong>en</strong>couragem<strong>en</strong>ts, son<br />

ai<strong>de</strong> pour corriger mes lettres <strong>en</strong> français et pour sa g<strong>en</strong>tillesse. Je lui souhaite bon courage<br />

pour sa thèse qui vi<strong>en</strong>t <strong>de</strong> démarrer.<br />

Je remercie mes anci<strong>en</strong>s collègues <strong>de</strong> bureau p<strong>en</strong>dant presque un an: Laur<strong>en</strong>t Risser,<br />

Pauline Assemat, Romain Guibert au bout du couloir, bureau 110, où j’ai comm<strong>en</strong>cé ma<br />

thèse.<br />

VIII


Toute mon amitié à Yohan Davit (le grand chef), Stephanie Veran (Mme Tissoires<br />

spécialiste <strong>de</strong>s mots fléchés ), Alexandre (le Grand) Lapène, Flor<strong>en</strong>t H<strong>en</strong>on (avec ou sans<br />

sabre chinois), Vinc<strong>en</strong>t Sarrot (champion <strong>de</strong>s chiffres et <strong>de</strong>s lettres), Yunli Wang<br />

(championne <strong>de</strong> rallye), Clém<strong>en</strong>t Louriou (dominateur d’informatique et d’acquisition <strong>de</strong>s<br />

données), les inséparables : Fabi<strong>en</strong> Chauvet + Ian Billanou, Dominique Courret (passionné<br />

<strong>de</strong> poissons), Bilal Elhajar (champion <strong>de</strong> t<strong>en</strong>nis), Arnaud Pujol (fameux ciné-man du<br />

groupe), Faiza Hidri, Sol<strong>en</strong>n Cotel, Haishan Luo, Hassane Fatmi, Karine Spielmann<br />

(championne <strong>de</strong> ping pong), Mehdi Rebai, Dami<strong>en</strong> Ch<strong>en</strong>u. Je les remercie pour leur amitié<br />

et pour leur souti<strong>en</strong> moral, avec eux j’ai vécu <strong>de</strong>s mom<strong>en</strong>ts inoubliables plein d’amitié<br />

avec ambiance et humour à coté du travail. Je remercie aussi tous les responsables et les<br />

membres <strong>de</strong> la fameuse pause café du groupe. Merci à tous, sans eux cette av<strong>en</strong>ture aurait<br />

sûrem<strong>en</strong>t été moins plaisante.<br />

Souv<strong>en</strong>t, parler dans sa langue maternelle ça ai<strong>de</strong> à oublier la nostalgie du pays ; je<br />

remercie donc Hossein Fadaei et sa femme qui ont organisé quelques randonnées durant<br />

ces années.<br />

Je suis très fier d’avoir appris la langue française, je remercie beaucoup mes<br />

professeurs <strong>de</strong> l’Alliance Française <strong>de</strong> Toulouse <strong>en</strong> particulier Sébasti<strong>en</strong> Palusci et Lucie<br />

Pépin. Grâce à Lucie j’ai beaucoup avancé <strong>en</strong> communication orale, je l’<strong>en</strong> remercie<br />

beaucoup. P<strong>en</strong>dant cette pério<strong>de</strong>, à l’Alliance Française <strong>de</strong> Toulouse, j’ai trouvé <strong>de</strong>s amis<br />

<strong>de</strong> tous les coins du mon<strong>de</strong>. Ils sont très nombreux et g<strong>en</strong>tils. Je remercie particulièrem<strong>en</strong>t<br />

Luciano Xavier, Isaac Suarez, Pavel Dub, Laia Moret Gabarro, Zaira Arellano, Fernando<br />

Maestre, Paula Margaretic, Azuc<strong>en</strong>a Castinera, Alan Llamas qui sont restés fidèles.<br />

Au cours <strong>de</strong> l’été 2009 j’ai participé à une école d’été sur la modélisation <strong>de</strong>s<br />

réservoirs pétroliers à l’université technique du Danemark (DTU) <strong>de</strong> Lyngby ; c’était un<br />

grand honneur pour moi <strong>de</strong> r<strong>en</strong>contrer Alexan<strong>de</strong>r Shapiro et ses collègues du départem<strong>en</strong>t<br />

<strong>de</strong> génie chimique et biochimique. Je remercie égalem<strong>en</strong>t Osvaldo Chiavone, Negar<br />

Sa<strong>de</strong>gh et Yok Pongthunya pour leur amitié p<strong>en</strong>dant cette pério<strong>de</strong>.<br />

Je remercie mes anci<strong>en</strong>s amis et mes anci<strong>en</strong>s collègues <strong>de</strong> l’université <strong>de</strong> Shahrood, je<br />

voulais leur dire que même si la distance nous sépare physiquem<strong>en</strong>t, l’esprit d’amitié est<br />

toujours resté <strong>en</strong>tre nous et je ne vous oublierai jamais.<br />

Enfin, je ti<strong>en</strong>s à remercier du fond du coeur mes par<strong>en</strong>ts et ma famille pour les<br />

<strong>en</strong>couragem<strong>en</strong>ts et le souti<strong>en</strong> qu’ils m’ont apporté tout au long du parcours qui m’a m<strong>en</strong>é<br />

jusqu’ici.<br />

« Be paian amad in daftar hekaiat hamch<strong>en</strong>an baghist !»<br />

(Ce cahier se termine, mais l’histoire continue !)<br />

IX


Table of Cont<strong>en</strong>ts<br />

1. G<strong>en</strong>eral Introduction ..............................................................2<br />

1.1 Industrial interest of Soret effect ........................................................... 4<br />

1.2 Theoretical Direct numerical solution (DNS) ....................................... 6<br />

1.3 Theoretical upscaling methods .............................................................. 6<br />

1.3.1 Multi-scale, hierarchical system .............................................................. 6<br />

1.3.2 Upscaling tools for porous media ............................................................ 9<br />

1.4 Experim<strong>en</strong>tal methods ......................................................................... 10<br />

1.4.1 Two-bulb method................................................................................... 10<br />

1.4.2 The Thermogravitational Column.......................................................... 12<br />

1.4.3 Thermal Field-Flow Fractionation (ThFFF) .......................................... 13<br />

1.4.4 Forced Rayleigh-Scattering Technique.................................................. 13<br />

1.4.5 The single-beam Z-scan or thermal l<strong>en</strong>s technique ............................... 14<br />

1.5 Conc<strong>en</strong>tration measurem<strong>en</strong>t ................................................................ 14<br />

1.5.1 From the variation of thermal conductivity ........................................... 15<br />

1.5.2 From the variation of viscosity .............................................................. 16<br />

1.5.3 Gas Chromatography (GC) .................................................................... 16<br />

1.5.4 Analysis by mass spectrometer.............................................................. 18<br />

1.6 Conclusion ........................................................................................... 19<br />

2. Theoretical predictions of the effective diffusion and<br />

thermal diffusion coeffici<strong>en</strong>ts in porous media ..........................21<br />

2.1 Introduction.......................................................................................... 25<br />

2.2 Governing microscopic equation......................................................... 27<br />

2.3 Volume averaging method................................................................... 29<br />

2.4 Darcy’s law .......................................................................................... 31<br />

2.4.1 Brinkman term ....................................................................................... 31<br />

2.4.2 No-linear case ........................................................................................ 32<br />

2.4.3 Low permeability correction.................................................................. 33<br />

XI


2.5 Transi<strong>en</strong>t conduction and convection heat transport ........................... 34<br />

2.5.1 One equation local thermal equilibrium ................................................ 36<br />

2.5.2 Two equation mo<strong>de</strong>l............................................................................... 48<br />

2.5.3 Non-equilibrium one-equation mo<strong>de</strong>l.................................................... 49<br />

2.6 Transi<strong>en</strong>t diffusion and convection mass transport ............................. 51<br />

2.6.1 Local closure problem............................................................................ 53<br />

2.6.2 Closed form............................................................................................ 56<br />

2.6.3 Non thermal equilibrium mo<strong>de</strong>l............................................................. 57<br />

2.7 Results.................................................................................................. 59<br />

2.7.1 Non-conductive solid-phase ( k ≈ 0 ) .................................................... 60<br />

σ<br />

2.7.2 Conductive solid-phase ( k ≠ 0)............................................................<br />

67<br />

σ<br />

2.7.3 Solid-solid contact effect ....................................................................... 71<br />

2.8 Conclusion ........................................................................................... 76<br />

3. Microscopic simulation and validation................................78<br />

3.1 Microscopic geometry and boundary conditions ................................ 79<br />

3.2 Non-conductive solid-phase ( k ≈ 0)...................................................<br />

80<br />

3.2.1 Pure diffusion ( 0, k ≈ 0)<br />

≈ σ<br />

σ<br />

Pe ................................................................ 80<br />

3.2.2 Diffusion and convection ( 0, k ≈ 0)<br />

Pe ............................................. 83<br />

≠ σ<br />

3.3 Conductive solid-phase ( k ≠ 0)<br />

.......................................................... 85<br />

3.3.1 Pure diffusion ( 0, k ≠ 0)<br />

σ<br />

Pe .............................................................. 85<br />

≈ σ<br />

3.3.2 Diffusion and convection ( 0, k ≠ 0)<br />

Pe ............................................. 92<br />

≠ σ<br />

3.4 Conclusion ........................................................................................... 97<br />

4. A new experim<strong>en</strong>tal setup to <strong>de</strong>termine the effective<br />

coeffici<strong>en</strong>ts .....................................................................................99<br />

4.1 Introduction........................................................................................ 102<br />

4.2 Experim<strong>en</strong>tal setup ............................................................................ 103<br />

4.2.1 Diffusion in a two-bulb cell ................................................................. 106<br />

4.2.2 Two-bulb apparatus <strong>en</strong>d correction ..................................................... 109<br />

XII


4.2.3 Thermal diffusion in a two-bulb cell ................................................... 110<br />

4.2.4 A transi<strong>en</strong>t-state method for thermal diffusion processes ................... 111<br />

4.3 Experim<strong>en</strong>tal setup for porous media................................................113<br />

4.4 Results................................................................................................ 113<br />

4.4.1 Katharometer calibration...................................................................... 113<br />

4.4.2 Diffusion coeffici<strong>en</strong>t ............................................................................ 115<br />

4.4.3 Effective diffusion coeffici<strong>en</strong>t in porous media................................... 117<br />

4.4.4 Free fluid and effective thermal diffusion coeffici<strong>en</strong>t ......................... 121<br />

4.4.5 Effect of solid thermal conductivity on thermal diffusion................... 127<br />

4.4.6 Effect of solid thermal connectivity on thermal diffusion................... 130<br />

4.4.7 Effect of tortuosity on diffusion and thermal diffusion coeffici<strong>en</strong>ts ... 132<br />

4.5 Discussion and comparison with theory............................................134<br />

4.6 Conclusion ......................................................................................... 137<br />

5. G<strong>en</strong>eral conclusions and perspectives................................139<br />

XIII


List of tables<br />

Table 1-1. Flux-force coupling betwe<strong>en</strong> heat and mass ........................................................ 5<br />

Table 2-1. Objectives of each or<strong>de</strong>r of mom<strong>en</strong>tum analysis ............................................... 49<br />

Table 4-1. Thermal conductivity and corresponding katharometer reading for some gases at<br />

atmospheric pressure and T=300°K................................................................................... 105<br />

Table 4-2. The properties of CO2, N2 and He required to calculate kmix<br />

XIV<br />

(T=300 °C, P=1<br />

atm.)................................................................................................................................... 115<br />

Table 4-3. Molecular weight and L<strong>en</strong>nard-Jones parameters necessary to estimate diffusion<br />

coeffici<strong>en</strong>t ......................................................................................................................... 117<br />

Table 4-4. Estimation of diffusion coeffici<strong>en</strong>ts for binary gas mixtures He-CO2 and He-N2<br />

at temperatures 300, 350 and T = 323.<br />

7 °K, pressure 1 bar.............................................. 117<br />

Table 4-5. Measured diffusion coeffici<strong>en</strong>t for He-N2 and differ<strong>en</strong>t media ...................... 120<br />

Table 4-6. Measured diffusion coeffici<strong>en</strong>t for He-CO2 and differ<strong>en</strong>t medium ................ 121<br />

Table 4-7. Measured thermal diffusion and diffusion coeffici<strong>en</strong>t for He-N2 and for differ<strong>en</strong>t<br />

media ................................................................................................................................. 124<br />

Table 4-8. Measured diffusion coeffici<strong>en</strong>t and thermal diffusion coeffici<strong>en</strong>t for He-CO2<br />

and for differ<strong>en</strong>t media ...................................................................................................... 125<br />

Table 4-9. Measured diffusion coeffici<strong>en</strong>t and thermal diffusion coeffici<strong>en</strong>t for He-N2 and<br />

differ<strong>en</strong>t media................................................................................................................... 127<br />

Table 4-10. The solid (spheres) and fluid mixture physical properties (T=300 K) .......... 128<br />

Table 4-11. The solid (spheres) and fluid mixture physical properties (T=300 K) .......... 131<br />

Table 4-12. Porous medium tortuosity coeffici<strong>en</strong>ts ......................................................... 133


List of figures<br />

Fig. 1-1 Example of a multi-scale system ........................................................................... 7<br />

Fig. 1-2. A schematic diagram of the two-bulb apparatus used to <strong>de</strong>termine the thermal<br />

diffusion factors for binary gas mixtures ............................................................................ 11<br />

Fig. 1-3. Principle of Thermogravitational Cell with a horizontal temperature gradi<strong>en</strong>t... 12<br />

Fig. 1-4. Principle of Thermal Field-Flow Fractionation (ThFFF) .................................... 13<br />

Fig. 1-5. Principle of forced Rayleigh scattering ............................................................... 14<br />

Fig. 1-6. Diagram showing vertical section of the katharometer ...................................... 15<br />

Fig. 1-7. Schematics of a Gas Chromatograph Flame Ionization Detector (GC-FID)...... 17<br />

Fig. 1-8. Schematics of a Gas Chromatograph Electron Capture Detector (GC-ECD) ... 17<br />

Fig. 1-9. Schematics of a simple mass spectrometer.......................................................... 18<br />

Fig. 2-1. Problem configuration ......................................................................................... 28<br />

Fig. 2-2. Normalized temperature versus position, for three differ<strong>en</strong>t times (triangle, Direct<br />

β<br />

Numerical Simulation= ( T −T<br />

) ( T −T<br />

)<br />

σ<br />

= ( T −T<br />

) ( T −T<br />

)<br />

C<br />

H<br />

C<br />

β<br />

C<br />

H<br />

C<br />

; circles, Direct Numerical Simulation<br />

; solid line, Local-equilibrium mo<strong>de</strong>l= ( T T ) ( T −T<br />

)<br />

σ C H C<br />

XV<br />

− ...................... 44<br />

Fig. 2-3. Normalized temperature versus position, for three differ<strong>en</strong>t times (triangle, Direct<br />

β<br />

Numerical Simulation= ( T −T<br />

) ( T −T<br />

)<br />

σ<br />

= ( T −T<br />

) ( T −T<br />

)<br />

β<br />

C<br />

H<br />

C<br />

; circles, Direct Numerical Simulation<br />

; solid line, Local-equilibrium mo<strong>de</strong>l= ( ) ( )<br />

σ C H C<br />

T TC<br />

TH<br />

−TC<br />

− ...................... 46<br />

Fig. 2-4. Chang’s unit cell .................................................................................................. 55<br />

Fig. 2-5. Spatially periodic arrangem<strong>en</strong>t of the phases ...................................................... 59<br />

Fig. 2-6. Repres<strong>en</strong>tative unit cell (εβ=0.8).......................................................................... 60<br />

Fig. 2-7. Effective diffusion, thermal diffusion and thermal conductivity coeffici<strong>en</strong>ts at<br />

Pe=0..................................................................................................................................... 62<br />

Fig. 2-8. Effective, longitudinal coeffici<strong>en</strong>ts as a function of Péclet number ( k ≈ 0 and<br />

ε 0.<br />

8 ): (a) mass dispersion , (b) thermal dispersion , (c) thermal diffusion and (d) Soret<br />

=<br />

β<br />

number................................................................................................................................. 65<br />

Fig. 2-9. Comparison of closure variables<br />

b and<br />

Sβ<br />

x<br />

b for εβ=0.8 ............................... 66<br />

Fig. 2-10. The influ<strong>en</strong>ce of conductivity ratio (κ ) on (a) effective, longitudinal thermal<br />

conductivity and (b) effective thermal diffusion coeffici<strong>en</strong>ts (εβ=0.8) ............................... 68<br />

Tβ<br />

x<br />

σ


Fig. 2-11. Comparison of closure variables fields b Tβ<br />

and b Sβ<br />

for differ<strong>en</strong>t thermal<br />

conductivity ratio ( )<br />

κ at pure diffusion ( 0 & ε = 0.<br />

8)<br />

Pe ................................................... 69<br />

= β<br />

Fig. 2-12. Comparison of closure variables fields<br />

conductivity ratio ( )<br />

XVI<br />

b and<br />

Tβ<br />

x<br />

κ at convective regime ( 14 & ε = 0.<br />

8)<br />

= β<br />

b for differ<strong>en</strong>t thermal<br />

Sβ<br />

x<br />

Pe ........................................... 70<br />

Fig. 2-13. The influ<strong>en</strong>ce of conductivity ratio (κ ) on the effective coeffici<strong>en</strong>ts by<br />

resolution of the closure problem in a Chang’s unit cell (εβ=0.8 , Pe=0)............................ 71<br />

Fig. 2-14. Spatially periodic mo<strong>de</strong>l for solid-solid contact ................................................ 72<br />

Fig. 2-15. Effective thermal conductivity for (a) non-touching particles, a/d=0 (b) touching<br />

particles, a/d=0.002, (εβ=0.36, Pe=0) .................................................................................. 72<br />

Fig. 2-16. Spatially periodic unit cell to solve the thermal diffusion closure problem with<br />

solid-solid connections a/d=0.002, (εβ=0.36, Pe=0)............................................................ 73<br />

Fig. 2-17. Effective thermal conductivity and thermal diffusion coeffici<strong>en</strong>t for touching<br />

particles, a/d=0.002, εβ=0.36, Pe=0..................................................................................... 74<br />

Fig. 2-18. Comparison of closure variables fields b Tβ<br />

and b Sβ<br />

wh<strong>en</strong> the solid phase is<br />

continue, for differ<strong>en</strong>t thermal conductivity ratio ( κ ) at pure diffusion................................ 75<br />

Fig. 2-19. Effective thermal conductivity and thermal diffusion coeffici<strong>en</strong>t for touching<br />

particles, a/d=0.002, εβ=0.36 ............................................................................................... 76<br />

Fig. 3-1. Schematic of a spatially periodic porous medium ( T H : Hot Temperature and T C :<br />

Cold Temperature)............................................................................................................... 79<br />

Fig. 3-2. Comparison betwe<strong>en</strong> theoretical and numerical results at diffusive regime and<br />

κ=0, (a) time evolution of the conc<strong>en</strong>tration at x = 15 and (b and c) instantaneous<br />

temperature and conc<strong>en</strong>tration field .................................................................................... 82<br />

Fig. 3-3. Comparison betwe<strong>en</strong> theoretical and numerical results, κ=0 and Pe=1, (a and b)<br />

instantaneous temperature and conc<strong>en</strong>tration field, (c) time evolution of the conc<strong>en</strong>tration<br />

at x = 0.5, 7.5 and 13.5 ....................................................................................................... 84<br />

Fig. 3-4. Influ<strong>en</strong>ce of the thermal conductivity ratio on the temperature and conc<strong>en</strong>tration<br />

fields .................................................................................................................................... 86<br />

Fig. 3-5. (a) Temperature and (b) conc<strong>en</strong>tration profiles for differ<strong>en</strong>t conductivity ratio 87<br />

Fig. 3-6. Temporal evolution of the separation profiles for differ<strong>en</strong>t thermal conductivity<br />

ratio...................................................................................................................................... 88<br />

Fig. 3-7. Comparison betwe<strong>en</strong> theoretical and numerical results at diffusive regime and<br />

κ=10, temporal evolution of (a) temperature and (b) conc<strong>en</strong>tration profiles ...................... 89


Fig. 3-8. Effect of thermal conductivity ratio at diffusive regime on (a and b)<br />

instantaneous temperature and conc<strong>en</strong>tration field at t=10 and (b) time evolution of the<br />

conc<strong>en</strong>tration at x = 15 ....................................................................................................... 91<br />

Fig. 3-9. Comparison betwe<strong>en</strong> theoretical and numerical results, κ=10 and Pe=1, (a) time<br />

evolution of the conc<strong>en</strong>tration at x = 0.5, 7.5 and 13.5 (b and c) instantaneous temperature<br />

and conc<strong>en</strong>tration field ........................................................................................................ 93<br />

Fig. 3-10. Influ<strong>en</strong>ce of Péclet number on steady-state (a) temperature and (b)<br />

conc<strong>en</strong>tration profiles (κ=10) .............................................................................................. 94<br />

Fig. 3-11. Influ<strong>en</strong>ce of Péclet number on steady-state conc<strong>en</strong>tration at the exit (κ=10).... 95<br />

Fig. 3-12. Influ<strong>en</strong>ce of (a) separation factor and (b) conductivity ratio on pick point of the<br />

conc<strong>en</strong>tration profile............................................................................................................ 96<br />

Fig. 4-1. Sketch of the two-bulb experim<strong>en</strong>tal set-up used for the diffusion and thermal<br />

diffusion tests..................................................................................................................... 104<br />

Fig. 4-2. Dim<strong>en</strong>sions of the <strong>de</strong>signed two-bulb apparatus used in this study .................. 104<br />

Fig. 4-3. Katharometer used in this study (CATARC MP – R) ....................................... 105<br />

Fig. 4-4. A schematic of katharometer connection to the bulb......................................... 106<br />

Fig. 4-5. Two-bulb apparatus ........................................................................................... 106<br />

Fig. 4-6. Katharometer calibration curve with related estimation of thermal conductivity<br />

values for the system He-CO2 ........................................................................................... 114<br />

Fig. 4-7. Solute transport process in porous media .......................................................... 115<br />

Fig. 4-8. Cylindrical samples filled with glass sphere..................................................... 118<br />

Fig. 4-9. X-ray tomography <strong>de</strong>vice (Skyscan 1174 type) used in this study.................... 119<br />

Fig. 4-10. Section images of the tube (inner diameter d = 0.<br />

795cm)<br />

filled by differ<strong>en</strong>t<br />

materials obtained by an X-ray tomography <strong>de</strong>vice (Skyscan 1174 type)........................ 119<br />

Fig. 4-11. Composition-time history in two-bulb diffusion cell for He-N2 system for<br />

differ<strong>en</strong>t medium. ( 300K<br />

T C<br />

0<br />

= and 100%<br />

c )..................................................................... 120<br />

1 = b<br />

Fig. 4-12. Composition-time history in two-bulb diffusion cell for He-CO2 system for<br />

0<br />

differ<strong>en</strong>t medium ( T = 300K<br />

and 100%<br />

c )....................................................................... 121<br />

1 b =<br />

Fig. 4-13. Schematic diagram of two bulb a) diffusion and b) thermal diffusion processes<br />

........................................................................................................................................... 122<br />

Fig. 4-14. Composition-time history in two-bulb thermal diffusion cell for He-N2 binary<br />

mixture for differ<strong>en</strong>t media ( T = 50K<br />

0<br />

Δ , T = 323.<br />

7K<br />

and 50%<br />

XVII<br />

c ) .................................... 124<br />

1 b =


Fig. 4-15. Composition-time history in two-bulb thermal diffusion cell for He-CO2 binary<br />

mixture for differ<strong>en</strong>t media .( T = 50K<br />

0<br />

Δ , T = 323.<br />

7K<br />

and 50%<br />

XVIII<br />

c ) ................................... 125<br />

Fig. 4-16. New experim<strong>en</strong>tal thermal diffusion setup without the valve betwe<strong>en</strong> the two<br />

bulbs .................................................................................................................................. 126<br />

Fig. 4-17. Composition-time history in two-bulb thermal diffusion cell for He-N2 binary<br />

mixture for differ<strong>en</strong>t media ( T = 50K<br />

1 b =<br />

0<br />

Δ , T = 323.<br />

7K<br />

and 61.<br />

25%<br />

c ) ................................ 127<br />

Fig. 4-18. Cylindrical samples filled with differ<strong>en</strong>t materials (H: stainless steal, G: glass<br />

spheres and ε=42.5) ........................................................................................................... 128<br />

Fig. 4-19. Katharometer reading time history in two-bulb thermal diffusion cell for He-<br />

CO2 binary mixture for porous media having differ<strong>en</strong>t thermal conductivity (3 samples of<br />

stainless steal and 3 samples of glass spheres) ( T = 50K<br />

1 b =<br />

0<br />

Δ , T = 323.<br />

7K<br />

and 50%<br />

c )....... 129<br />

Fig. 4-20. Cylindrical samples filled with differ<strong>en</strong>t materials (A: glass spheres, B:<br />

aluminium spheres and ε=0.56)......................................................................................... 130<br />

Fig. 4-21. Katharometer time history in two-bulb thermal diffusion cell for He-CO2 binary<br />

mixture for porous media ma<strong>de</strong> of differ<strong>en</strong>t thermal conductivity (aluminum and glass<br />

spheres) ( Δ T = 50K<br />

, T 323.<br />

7K<br />

0<br />

= and 50%<br />

1 b =<br />

1 = b<br />

c ) .................................................................. 131<br />

Fig. 4-22. Definition of tortuosity coeffici<strong>en</strong>t in porous media, L= straight line and L’=<br />

real path l<strong>en</strong>gth .................................................................................................................. 132<br />

Fig. 4-23. Cylindrical samples filled with differ<strong>en</strong>t materials producing differ<strong>en</strong>t<br />

tortuosity but the same porosity ε=66% (E: cylindrical material and F: glass wool)........ 133<br />

Fig. 4-24. Composition time history in two-bulb thermal diffusion cell for He-CO2 binary<br />

mixture in porous media ma<strong>de</strong> of the same porosity (ε=66% ) but differ<strong>en</strong>t tortuosity<br />

(cylindrical materials and glass wool) ( T = 50K<br />

0<br />

Δ , T = 323.<br />

7K<br />

and 50%<br />

c ) ................... 134<br />

Fig. 4-25. Comparison of experim<strong>en</strong>tal effective diffusion coeffici<strong>en</strong>t data with the<br />

theoretical one obtained from volume averaging technique for differ<strong>en</strong>t porosity and a<br />

specific unit cell................................................................................................................. 135<br />

Fig. 4-26. Comparison of experim<strong>en</strong>tal effective thermal diffusion coeffici<strong>en</strong>t data with<br />

theoretical one obtained from volume averaging technique for differ<strong>en</strong>t porosity and a<br />

specific unit cell................................................................................................................. 136<br />

Fig. 4-27. Comparison of the experim<strong>en</strong>tal thermal diffusion ratio data with theoretical<br />

one obtained from volume averaging technique for differ<strong>en</strong>t porosity and a specific unit<br />

cell ..................................................................................................................................... 136<br />

1 b =


Fig. 5-1. 3D geometry of the closure problem with particle-particle touching ma<strong>de</strong> with<br />

COMSOL Multiphysics..................................................................................................... 141<br />

Fig. 5-2. Discrepancy betwe<strong>en</strong> numerical results and experim<strong>en</strong>tal measurem<strong>en</strong>ts in a<br />

packed thermo- gravitational cell ..................................................................................... 142<br />

Fig. 5-3. Proposition of experim<strong>en</strong>tal setup for convective regime ................................. 143<br />

XIX


Chapter 1<br />

G<strong>en</strong>eral Introduction


1. G<strong>en</strong>eral Introduction<br />

The Ludwig-Soret effect, also known as thermal diffusion (or thermal diffusion and also<br />

thermo-migration), is a classic example of coupled heat and mass transport in which the<br />

motion of the particles in a fluid mixture is driv<strong>en</strong> by a heat flux coming from a thermal<br />

gradi<strong>en</strong>t. G<strong>en</strong>erally, heaviest particle moves from hot to cold, but the reverse is also se<strong>en</strong><br />

un<strong>de</strong>r some conditions. The Soret effect has be<strong>en</strong> studied for about 150 years with more<br />

active periods following economic interests (separation of isotopes in the 30s, petroleum<br />

<strong>en</strong>gineering in the 90s ...). Many researchers have <strong>de</strong>veloped differ<strong>en</strong>t techniques to<br />

measure this effect and <strong>de</strong>duced theories to explain it. However, because of the complexity<br />

of this coupled ph<strong>en</strong>om<strong>en</strong>on, only rec<strong>en</strong>tly, there has be<strong>en</strong> an agreem<strong>en</strong>t on the values of<br />

the thermal diffusion coeffici<strong>en</strong>ts measured by differ<strong>en</strong>t techniques. Theoretically, there<br />

exists a rigorous approach based on the kinetic gas theory which explains the thermal<br />

diffusion effect for binary and multi-compon<strong>en</strong>t i<strong>de</strong>al gas mixtures. For liquids, the<br />

theories <strong>de</strong>veloped are not <strong>en</strong>ough accurate and there is still a lack of un<strong>de</strong>rstanding on the<br />

basis of the effect for these mixtures. The situation becomes ev<strong>en</strong> more complicated wh<strong>en</strong><br />

consi<strong>de</strong>ring porous media. Fluid and flow problems in porous media have attracted the<br />

att<strong>en</strong>tion of industrialists, <strong>en</strong>gineers and sci<strong>en</strong>tists from varying disciplines, such as<br />

chemical, <strong>en</strong>vironm<strong>en</strong>tal, and mechanical <strong>en</strong>gineering, geothermal physics and food<br />

sci<strong>en</strong>ce. The main goal of the pres<strong>en</strong>t thesis is to un<strong>de</strong>rstand this complexity in porous<br />

media wh<strong>en</strong> there is a coupling betwe<strong>en</strong> heat and mass transfer. The main objective is to<br />

study if the effective thermal diffusion <strong>de</strong>p<strong>en</strong>ds on the following<br />

• the void fraction of the phases and the structure of the solid matrix, i.e., the ext<strong>en</strong>t<br />

of the continuity of the solid phase,<br />

• the thermal conductivity of each phase, i.e., the relative magnitu<strong>de</strong> of thermal<br />

conductivity ratio,<br />

• the contact betwe<strong>en</strong> the no-consolidated particles, i.e., the solid surface coatings,<br />

• the fluid velocity, i.e., dispersion and free convection in pore spaces.<br />

The background and main goal of this thesis is pres<strong>en</strong>ted in this chapter.<br />

In chapter 2 we pres<strong>en</strong>t a theoretical approach based on the volume averaging method to<br />

<strong>de</strong>termine the effective transport coeffici<strong>en</strong>ts in porous media. In this part, we are<br />

2


interested in the upscaling of mass and <strong>en</strong>ergy coupled conservation equations of each<br />

compon<strong>en</strong>t of the mixture.<br />

Chapter 3 pres<strong>en</strong>ts a validation of the proposed theory by comparing the predicted<br />

behavior to results obtained from a direct pore-scale simulation.<br />

In chapter 4, coeffici<strong>en</strong>ts of diffusion and thermal diffusion are measured directly using<br />

specially <strong>de</strong>signed two-bulb method, and differ<strong>en</strong>t synthetic porous media with differ<strong>en</strong>t<br />

properties.<br />

Finally, in chapter 5, conclusions and suggestions for future work are pres<strong>en</strong>ted.<br />

Introduction générale <strong>en</strong> français<br />

L’effet <strong>de</strong> Ludwig-Soret, égalem<strong>en</strong>t connu sous le nom <strong>de</strong> thermal diffusion (ou thermomigration),<br />

est un exemple classique <strong>de</strong> phénomène couplé <strong>de</strong> transport <strong>de</strong> chaleur et<br />

matière dans lequel le mouvem<strong>en</strong>t molécules (ou <strong>de</strong>s particules) dans un mélange flui<strong>de</strong> est<br />

produit par un flux <strong>de</strong> chaleur dérivant d’un gradi<strong>en</strong>t thermique. En général, la particule la<br />

plus lour<strong>de</strong> se dirige vers la région plus froi<strong>de</strong>, mais l'inverse est égalem<strong>en</strong>t possible sous<br />

certaines conditions. L'effet Soret est étudié <strong>de</strong>puis prés <strong>de</strong> 150 ans avec <strong>de</strong>s pério<strong>de</strong>s plus<br />

actives suivant les intérêts économiques (séparations d’isotopes dans les années 30, génie<br />

pétrolier dans les années 90 …). Différ<strong>en</strong>tes techniques ont été mises aux points pour<br />

mesurer cet effet et développer les théories pour l'expliquer. Toutefois, <strong>en</strong> raison <strong>de</strong> la<br />

complexité <strong>de</strong> ce phénomène couplé, ce n’est que récemm<strong>en</strong>t qu’il y a eu un accord sur les<br />

valeurs <strong>de</strong>s coeffici<strong>en</strong>ts <strong>de</strong> thermal diffusion mesurées par <strong>de</strong>s techniques différ<strong>en</strong>tes.<br />

Théoriquem<strong>en</strong>t, il existe une approche rigoureuse basée sur la théorie cinétique <strong>de</strong>s gaz qui<br />

explique l'effet <strong>de</strong> thermal diffusion pour les mélanges binaires et multi-composants <strong>de</strong> gaz<br />

parfaits. Pour les liqui<strong>de</strong>s, les théories développées ne sont pas assez précises et il y a<br />

toujours un manque <strong>de</strong> compréh<strong>en</strong>sion sur les fon<strong>de</strong>m<strong>en</strong>ts <strong>de</strong> cet effet. La situation <strong>de</strong>vi<strong>en</strong>t<br />

<strong>en</strong>core plus compliquée lorsque l'on considère cet effet <strong>en</strong> milieux poreux. <strong>Les</strong> problèmes<br />

d'écoulem<strong>en</strong>t <strong>de</strong> flui<strong>de</strong>s mutlticonstituants <strong>en</strong> milieu poreux <strong>en</strong> prés<strong>en</strong>ce <strong>de</strong> gradi<strong>en</strong>ts<br />

thermiques ont attiré l'att<strong>en</strong>tion <strong>de</strong>s industriels, <strong>de</strong>s ingénieurs et <strong>de</strong>s sci<strong>en</strong>tifiques <strong>de</strong><br />

différ<strong>en</strong>tes disciplines, telles que la chimie, l'<strong>en</strong>vironnem<strong>en</strong>t, le génie mécanique, la<br />

physique géothermique et sci<strong>en</strong>ce <strong>de</strong>s alim<strong>en</strong>ts. L'objectif principal <strong>de</strong> cette thèse est <strong>de</strong><br />

compr<strong>en</strong>dre cette complexité dans les milieux poreux lorsqu'il existe un couplage <strong>en</strong>tre les<br />

3


transferts <strong>de</strong> chaleur et <strong>de</strong> matière. L'objectif est d'étudier si la thermal diffusion effective<br />

dép<strong>en</strong>d <strong>de</strong> :<br />

• la fraction <strong>de</strong> vi<strong>de</strong> dans le milieu (porosité) et la structure <strong>de</strong> la matrice soli<strong>de</strong>, par<br />

exemple la continuité <strong>de</strong> la phase soli<strong>de</strong>,<br />

• la conductivité thermique <strong>de</strong> chaque phase, et <strong>en</strong> particulier la valeur du rapport<br />

<strong>de</strong>s conductivités thermiques,<br />

• le contact <strong>en</strong>tre les particules non-consolidées, et la forme générale <strong>de</strong> la surface<br />

d’échange <strong>de</strong> la matrice soli<strong>de</strong>,<br />

• la vitesse du flui<strong>de</strong>, c'est à dire la dispersion et la convection naturelle dans les<br />

espaces du pore.<br />

Le contexte et l'objectif principal <strong>de</strong> cette thèse sont prés<strong>en</strong>tés dans le chapitre 1.<br />

En chapitre 2, nous prés<strong>en</strong>tons une approche théorique basée sur la métho<strong>de</strong> <strong>de</strong> prise <strong>de</strong><br />

moy<strong>en</strong>ne volumique afin <strong>de</strong> déterminer les coeffici<strong>en</strong>ts <strong>de</strong> transport effectifs dans un<br />

milieu poreux. Dans cette partie, nous appliquons les techniques <strong>de</strong> changem<strong>en</strong>t d’échelle<br />

<strong>de</strong>s équations couplées <strong>de</strong> conservations <strong>de</strong> la matière et <strong>de</strong> l'énergie.<br />

Le chapitre 3 prés<strong>en</strong>tes une validation <strong>de</strong> la théorie proposée <strong>en</strong> comparant les résultats<br />

théoriques avec les résultats obt<strong>en</strong>us par simulation directe d’échelle du pore.<br />

En chapitre 4, les coeffici<strong>en</strong>ts <strong>de</strong> diffusion et <strong>de</strong> thermal diffusion sont mesurés<br />

directem<strong>en</strong>t <strong>en</strong> utilisant un dispositif expérim<strong>en</strong>tal à <strong>de</strong>ux bulbes, développé<br />

spécifiquem<strong>en</strong>t pour ce travail, et appliqué à différ<strong>en</strong>ts milieux poreux modèles réalisés<br />

dans différ<strong>en</strong>tes gammes <strong>de</strong> propriétés thermo-physique.<br />

Enfin, <strong>en</strong> chapitre 5, les conclusions et suggestions pour les travaux futurs sont prés<strong>en</strong>tées.<br />

1.1 Industrial interest of Soret effect<br />

In or<strong>de</strong>r to optimize production costs wh<strong>en</strong> extracting fluid field by producers, it is<br />

important to know precisely the distribution of differ<strong>en</strong>t species in the field. This<br />

distribution has g<strong>en</strong>erally be<strong>en</strong> g<strong>en</strong>erated over long formation period and separation has<br />

be<strong>en</strong> mainly influ<strong>en</strong>ced by the gravity and the distribution of pressure in the reservoir.<br />

Consi<strong>de</strong>rable methods have be<strong>en</strong> implem<strong>en</strong>ted in or<strong>de</strong>r to obtain reliable thermodynamic<br />

mo<strong>de</strong>ls, allow obtaining correctly the distribution of species in the reservoir. Since it is not<br />

4


possible to ignore the important vertical ext<strong>en</strong>sion of a giv<strong>en</strong> field, it is very possibly that<br />

this distribution is influ<strong>en</strong>ced by thermal diffusion and convection (gravity is one of the<br />

first compon<strong>en</strong>ts inclu<strong>de</strong>d in the mo<strong>de</strong>ls), but also by the geothermal gradi<strong>en</strong>t (natural<br />

temperature gradi<strong>en</strong>t of the earth).<br />

This gradi<strong>en</strong>t could be the cause of migration of species in a ph<strong>en</strong>om<strong>en</strong>on known as the<br />

Soret effect or thermal diffusion (more g<strong>en</strong>erally, the name thermal diffusion is used to<br />

<strong>de</strong>scribe this effect in a gas mixture; whereas Soret effect or Ludwig effect will be used in<br />

liquids). This is the creation of a conc<strong>en</strong>tration gradi<strong>en</strong>t of the chemical species by the<br />

pres<strong>en</strong>ce of a thermal gradi<strong>en</strong>t, i.e., the exist<strong>en</strong>ce of a thermal gradi<strong>en</strong>t is causing migration<br />

of species. This effect, discovered by C. Ludwig in 1856 [55] (and better exploited by C.<br />

Soret in 1880 [98]) is a particular ph<strong>en</strong>om<strong>en</strong>on since it is associated to coupled<br />

thermodynamic ph<strong>en</strong>om<strong>en</strong>a, i.e. a flux created by a force of differ<strong>en</strong>t nature (here a<br />

conc<strong>en</strong>tration gradi<strong>en</strong>t is induced by the pres<strong>en</strong>ce of a thermal gradi<strong>en</strong>t), Table 1-1<br />

summarizes the flux-force coupling effects betwe<strong>en</strong> heat and mass transfer.<br />

Table 1-1. Flux-force coupling betwe<strong>en</strong> heat and mass<br />

Flux\Force ∇ T<br />

∇ c<br />

Heat Fourier’s law of conduction Dufour effect<br />

Mass Soret effect Fick’s law of diffusion<br />

The study of these relations betwe<strong>en</strong> flux and forces of this type is called Thermodynamics<br />

of Linear Irreversible Processes [38]. The main characteristic quantity for thermal diffusion<br />

is a coeffici<strong>en</strong>t called Soret coeffici<strong>en</strong>t ( S T ). Many works have be<strong>en</strong> un<strong>de</strong>rtak<strong>en</strong> to<br />

<strong>de</strong>termine this quantity with differ<strong>en</strong>t approaches: experim<strong>en</strong>tal approaches (Soret<br />

Coeffici<strong>en</strong>ts in Cru<strong>de</strong> Oil un<strong>de</strong>r microgravity condition [35, 100], thermo-gravitational<br />

column) or theoretical approaches (molecular dynamics simulations [89, 34], multicompon<strong>en</strong>t<br />

numerical mo<strong>de</strong>ls [91]). Most of these research conclu<strong>de</strong>d that values obtained<br />

experim<strong>en</strong>tally are differ<strong>en</strong>t from the theoretical one. These differ<strong>en</strong>ces are mainly<br />

explained by the fact that the measurem<strong>en</strong>ts are technically simpler in a free medium<br />

(without the porous matrix), and the effects due to pore-scale velocity fluctuations or to<br />

differ<strong>en</strong>ces in thermal conductivity betwe<strong>en</strong> rock and liquid are th<strong>en</strong> not tak<strong>en</strong> into<br />

account. Failures in the thermo-gravitational mo<strong>de</strong>l based on the free fluid equations is a<br />

5


good example of the need to <strong>de</strong>termine a new mo<strong>de</strong>l for the ph<strong>en</strong>om<strong>en</strong>a of diffusion and<br />

thermal diffusion in porous media. There are several theoretical and experim<strong>en</strong>tal methods<br />

available to <strong>de</strong>termine the transport properties in porous media<br />

1.2 Theoretical Direct numerical solution (DNS)<br />

The direct numerical simulation of flows through porous formations is difficult due to the<br />

medium fine scale heterog<strong>en</strong>eity and also the complexity of dynamic systems. An accurate<br />

well-resolved computation oft<strong>en</strong> requires great amount of computer memory and CPU<br />

time, which can easily exceed the limit of today’s computer resources.<br />

Despite of this difficulty, the direct resolution of microscopic equation in porous media can<br />

be interesting for reasons of fundam<strong>en</strong>tal research, e.g., validation of macroscopic mo<strong>de</strong>ls<br />

(see for example [80] and [19]) as we have done in this study (Chapter 3). There are also<br />

many problems for which the upscaling processes are not possible or they are very difficult<br />

to achieve; therefore DNS can be used to resolve the problem in a simpler geometry<br />

problem on a volume containing a small number of pores.<br />

In practice, it is oft<strong>en</strong> suffici<strong>en</strong>t to predict the large scale solutions to certain accuracy.<br />

Therefore, alternative theoretical approaches have be<strong>en</strong> <strong>de</strong>veloped.<br />

1.3 Theoretical upscaling methods<br />

The un<strong>de</strong>rstanding and prediction of the behavior of the flow of multiphase or<br />

multicompon<strong>en</strong>t fluids through porous media are oft<strong>en</strong> strongly influ<strong>en</strong>ced by<br />

heterog<strong>en</strong>eities, either large-scale lithological discontinuities or quite localized ph<strong>en</strong>om<strong>en</strong>a<br />

[29]. Consi<strong>de</strong>rable information can be gained about the physics of multiphase flow of<br />

fluids through porous media via laboratory experim<strong>en</strong>ts and pore-scale mo<strong>de</strong>ls; however,<br />

the l<strong>en</strong>gth scales of these data are quite differ<strong>en</strong>t from those required from field-scale<br />

simulations. The pres<strong>en</strong>ce of heterog<strong>en</strong>eities in the medium also greatly complicates the<br />

flow. Therefore, we must un<strong>de</strong>rstand the effects of heterog<strong>en</strong>eities and coeffici<strong>en</strong>ts on<br />

differ<strong>en</strong>t l<strong>en</strong>gth scales.<br />

1.3.1 Multi-scale, hierarchical system<br />

Observation and mo<strong>de</strong>lling scales (Fig. 1-1) can be classified as<br />

• microscopic scale or pore scale,<br />

• macroscopic or Darcy scale, usually a few characteristic dim<strong>en</strong>sions of the pore,<br />

6


• mesoscopic or macroscopic scale heterog<strong>en</strong>eities of the porous medium, which<br />

correspond to variations in facies,<br />

• megascopic scale or scale of the aquifer, reservoir, etc.<br />

The physical <strong>de</strong>scription of the first two scales has be<strong>en</strong> the subject of many studies.<br />

Taking into account the effect of heterog<strong>en</strong>eity, poses many problems oft<strong>en</strong> unresolved<br />

wh<strong>en</strong> level <strong>de</strong>scription in the mo<strong>de</strong>l used is too large (e.g. a mesh numerical mo<strong>de</strong>l too<br />

large compared to heterog<strong>en</strong>eities).<br />

In a porous medium, the equations of continuum mechanics permit to <strong>de</strong>scribe the<br />

transport processes within the pores. For a large number of pores, the <strong>de</strong>tailed <strong>de</strong>scription<br />

of microscopic processes is g<strong>en</strong>erally impractical. It is therefore necessary to move from a<br />

microscopic <strong>de</strong>scription at the pore scale to a macroscopic <strong>de</strong>scription throughout a certain<br />

volume of porous medium including a large number of pores.<br />

In this section we <strong>de</strong>scribe briefly these differ<strong>en</strong>t scales and their influ<strong>en</strong>ces on the<br />

transport equations.<br />

β<br />

γ<br />

σ<br />

Microscopic<br />

or pore scale<br />

ω<br />

Macroscopic<br />

or Darcy scale<br />

Microscopic scale: β=water phase, σ=solid phase, γ=organic phase.<br />

7<br />

Megascopic<br />

or aquifer scale<br />

Macroscopic scale: η and ω are porous media of differ<strong>en</strong>t characteristics.<br />

Megascopic scale: here the aquifer contains two mesoscopic heterog<strong>en</strong>eities.<br />

Fig. 1-1 Example of a multi-scale system


I. Microscopic scale<br />

The microscopic <strong>de</strong>scription focuses on the behavior of a large number of molecules of the<br />

pres<strong>en</strong>t phases (e.g., organic phase and water phase shown in Fig. 1-1). The equations<br />

<strong>de</strong>scribing their transport are those of the continum mechanics. The flow is well <strong>de</strong>scribed<br />

by the following equations<br />

• Mass balance equations for all compon<strong>en</strong>ts in the consi<strong>de</strong>red phase. In these<br />

equations may appear, in addition to the accumulation, convection and diffusion<br />

terms, chemical reaction terms known as homog<strong>en</strong>eous chemical reaction as they<br />

take place within this phase<br />

• The Navier-Stokes equations <strong>de</strong>scribing the mom<strong>en</strong>tum balance<br />

• The equation of heat transfer if there are temperature gradi<strong>en</strong>ts in the system<br />

• Boundary conditions on interfaces with other phases which <strong>de</strong>p<strong>en</strong>d upon the<br />

physics of the problem.<br />

II. Macroscopic scale<br />

The direct resolution of microscopic equations on a volume containing a small numbers of<br />

pores is usually possible and interesting for reasons of fundam<strong>en</strong>tal research (e.g.<br />

validation of macroscopic mo<strong>de</strong>ls). However, it is usually impossible to solve these<br />

microscopic equations on a large volume. In practice, it must be obtained a macroscopic<br />

<strong>de</strong>scription repres<strong>en</strong>ting the effective behavior of the porous medium for a repres<strong>en</strong>tative<br />

elem<strong>en</strong>tary volume (REV) containing many pores. Many techniques have be<strong>en</strong> used to<br />

move from the pore scale to the REV scale [23]. Integration on the REV (called volume<br />

averaging technique) of the microscopic conservation equations allow obtaining<br />

macroscopic equations which are valid for average variables called macroscopic variables<br />

[7, 91].<br />

In the case of a homog<strong>en</strong>eous porous medium, the REV size can be characterized by a<br />

sphere whose diameter is about 30 times the average grain diameter [7]. The problems<br />

associated with upscaling from the microscopic scale to the macroscopic scale will be<br />

treated in the next chapter.<br />

At the macroscopic scale, the <strong>de</strong>scription of the flow of phases introduces new equations<br />

which are the transposition of the mass balance, mom<strong>en</strong>tum and <strong>en</strong>ergy microscopic<br />

equations. For example, the equation of Darcy is the mom<strong>en</strong>tum balance at the<br />

8


macroscopic scale. In these macroscopic equations appear effective properties, as the<br />

permeability in Darcy's law, the relative permeabilities and capillary pressure in the<br />

multiphase case, etc. These effective properties can be theoretically <strong>de</strong>duced from<br />

microscopic properties by using upscaling techniques. They are most oft<strong>en</strong> estimated from<br />

measurem<strong>en</strong>ts on a macroscopic scale. The direct measurem<strong>en</strong>t of these properties is not<br />

simple, because of heterog<strong>en</strong>eities of the medium.<br />

III. Mesoscopic and Megascopic scale<br />

The macroscopic properties are rarely the same at every point of the aquifer. Natural<br />

medium are in fact g<strong>en</strong>erally heterog<strong>en</strong>eous. It is sometimes possible to take into account<br />

the effect of these heterog<strong>en</strong>eities by solving the equations with a macroscopic mesh size<br />

smaller than the characteristic size of the heterog<strong>en</strong>eities. If this is not possible, the<br />

situation is similar to that already <strong>en</strong>countered in the transition betwe<strong>en</strong> microscopic and<br />

macroscopic scales: it must be established a valid <strong>de</strong>scription at the mesoscopic or<br />

megascopic.<br />

1.3.2 Upscaling tools for porous media<br />

In the macroscopic <strong>de</strong>scription of mass and heat transfer in porous media, the convectiondiffusion<br />

ph<strong>en</strong>om<strong>en</strong>a (or dispersion) are g<strong>en</strong>erally analyzed using an up-scaling method, in<br />

which the complicated local situation (transport by convection and diffusion at the pore<br />

scale) is finally <strong>de</strong>scribed at the macroscopic scale by effective t<strong>en</strong>sors [65]. To mo<strong>de</strong>l<br />

transport ph<strong>en</strong>om<strong>en</strong>a in porous media, several methods exist. These tools are listed below<br />

• integral transform methods,<br />

• fractional approaches,<br />

• homog<strong>en</strong>ization,<br />

• volume averaging technique<br />

• c<strong>en</strong>tral limit approaches,<br />

• Taylor–Aris–Br<strong>en</strong>ner (TAB) mom<strong>en</strong>t methods,<br />

• spectral integral approaches,<br />

• Fast Fourier transform (FFT) and Gre<strong>en</strong>s functions methods,<br />

• mixture and hybrid averaging/mixture approaches,<br />

9


• projection operator methods,<br />

• stationary stochastic convective type approaches,<br />

• and nonstationary stochastic convective type methods.<br />

The rea<strong>de</strong>r can look at [23] for a brief <strong>de</strong>scription of differ<strong>en</strong>t types of hierarchies and<br />

recomm<strong>en</strong><strong>de</strong>d tools which may be applied.<br />

Among others, the method of mom<strong>en</strong>ts [11], the volume averaging method [14] and the<br />

homog<strong>en</strong>ization method [61] are the most used techniques. In this study, we shall use the<br />

volume averaging method to obtain the macro-scale equations that <strong>de</strong>scribe thermal<br />

diffusion in a homog<strong>en</strong>eous porous medium [23]. It has be<strong>en</strong> ext<strong>en</strong>sively used to predict<br />

the effective transport properties for many processes including transport in heterog<strong>en</strong>eous<br />

porous media [83], two-phase flow [79], two-Phase inertial flow [53], reactive media [111,<br />

1], solute transport with adsorption [2] multi-compon<strong>en</strong>t mixtures [80].<br />

1.4 Experim<strong>en</strong>tal methods<br />

In this section, we pres<strong>en</strong>t a review of differ<strong>en</strong>t methods used for measuring the diffusion<br />

and thermal diffusion effect in gas. There is more than 150 years that the thermal diffusion<br />

effect was firstly observed by Ludwig. Along these years, researchers have <strong>de</strong>signed a<br />

wi<strong>de</strong> variety of setup for measuring this effect. Measuring thermal diffusion compared to<br />

diffusion and dispersion is not an easy task because this effect is usually very small and<br />

slow.<br />

In this section, the goal is not to explain all existing methods, but to <strong>de</strong>scribe briefly the<br />

methods most commonly used.<br />

1.4.1 Two-bulb method<br />

The two-bulb technique is the most wi<strong>de</strong>ly used method for <strong>de</strong>termining the diffusion<br />

coeffici<strong>en</strong>ts [114] and thermal diffusion [37] coeffici<strong>en</strong>ts of gases. The basic arrangem<strong>en</strong>t<br />

for a two-bulb cell consists of two chambers of relatively large volume joined by a smallvolume<br />

diffusion tube. Initially, the two chambers are filled with fluid mixtures of differ<strong>en</strong>t<br />

composition at the same pressure which are allowed to approach a uniform composition by<br />

means of diffusion through the tube.<br />

Experim<strong>en</strong>tal investigations of thermal diffusion have usually be<strong>en</strong> based on the<br />

<strong>de</strong>termination of the differ<strong>en</strong>ce in composition of two parts of a fluid mixture which are at<br />

10


differ<strong>en</strong>t temperatures. A temperature gradi<strong>en</strong>t is set up in the tube by bringing the bulbs to<br />

differ<strong>en</strong>t temperatures, uniform over each bulb. Provi<strong>de</strong>d the ratio of the bulbs volume is<br />

known, the separation can be found from the change in composition which occurs in one<br />

bulb only. A two-bulb apparatus used to <strong>de</strong>termine the thermal diffusion coeffici<strong>en</strong>ts is<br />

illustrated in Fig. 1-2. In this type of the two-bulb apparatus due to the large ratio in the<br />

volume of the two bulbs, almost all change in the gas mixture composition occurs in the<br />

lower bulb. In the literature, numerous measurem<strong>en</strong>ts were ma<strong>de</strong> in free medium in 50s<br />

and 60s (see for instance some series of measurem<strong>en</strong>ts which were done by Ibbs, 1921;<br />

Heath, 1941; van Itterbeek, 1947; Mason, 1962; Sax<strong>en</strong>a, 1966; Humphreys, 1970; Grew,<br />

1977; Shashkov, 1979 and Zhdanov, 1980).<br />

A: top bulb; B: bottom bulb; C: gas inlet valve; D: thermocouple; E: metal<br />

jacket; F: metal block; G and H: thermistor elem<strong>en</strong>ts and I: isolation valve<br />

Fig. 1-2. A schematic diagram of the two-bulb apparatus used to <strong>de</strong>termine the thermal diffusion factors for<br />

binary gas mixtures [95]<br />

11


1.4.2 The Thermogravitational Column<br />

Another method for measuring thermal diffusion coeffici<strong>en</strong>ts is the thermogravitational<br />

column which consists of two vertical plates separated by a narrow space un<strong>de</strong>r a<br />

horizontal [54] or vertical [30] thermal gradi<strong>en</strong>t. The principle is to use a thermal gradi<strong>en</strong>t<br />

to simultaneously produce a mass flux by thermal diffusion and a convection flux. Starting<br />

from a mixture of homog<strong>en</strong>eous composition, the coupling of the two transport<br />

mechanisms leads to a separation of the compon<strong>en</strong>ts. In most experim<strong>en</strong>tal <strong>de</strong>vices, the<br />

applied thermal gradi<strong>en</strong>t is horizontal and the final composition gradi<strong>en</strong>t is globally<br />

vertical. The separation rate in this system <strong>de</strong>fined as the conc<strong>en</strong>tration differ<strong>en</strong>ce betwe<strong>en</strong><br />

the top and the bottom cell. Thermogravitational column was <strong>de</strong>vised by Clusius and<br />

Dickel (1938). The ph<strong>en</strong>om<strong>en</strong>ology of thermogravitational transport was exposed by Furry<br />

et al. (1939), and was validated by many experim<strong>en</strong>ts. The optimal coupling betwe<strong>en</strong><br />

thermal diffusion and convection ratio (maximum separation) correspond to an optimal<br />

thickness of the cell in free fluid (less than one millimetre for usual liquids) and an optimal<br />

permeability in porous medium [56, 57]. The so called packed thermal diffusion cell (PTC)<br />

was <strong>de</strong>scribed and int<strong>en</strong>sively used to perform experim<strong>en</strong>ts on varieties of ionic and<br />

organic mixtures [54, 21, 66]. The separation in a thermogravitational column can be<br />

substantially increased by inclining the column [72]. Rec<strong>en</strong>tly, Mojtabi et al., 2003,<br />

showed that the vibrations can lead whether to increase or to <strong>de</strong>crease heat and mass<br />

transfers or <strong>de</strong>lay or accelerate the onset of convection [18].<br />

Cold Wall<br />

Thermal<br />

diffusion<br />

+<br />

Convection<br />

Fig. 1-3. Principle of Thermogravitational Cell with a horizontal temperature gradi<strong>en</strong>t<br />

12<br />

Hot Wall


1.4.3 Thermal Field-Flow Fractionation (ThFFF)<br />

Thermal field-flow fractionation (ThFFF) is a sub-technique of the FFF family that relies<br />

on a temperature gradi<strong>en</strong>t (create a thermal diffusion force) to characterize and separate<br />

polymers and particles. A schematic of the TFFF system is shown in Fig. 1-4. Separation<br />

of susp<strong>en</strong><strong>de</strong>d particles is typically performed in a solv<strong>en</strong>t carrier. Higher molecular weight<br />

particles react more to the thermal gradi<strong>en</strong>t and compact more tightly against the cold.<br />

Because of the parabolic velocity profile of the carrier, lower molecular weight will have a<br />

higher average velocity. The differ<strong>en</strong>ce in average velocity results in the spatial and<br />

temporal separation along the ThFFF channel. The TFFF system possesses unique<br />

characteristics making it more suitable for some separations than conv<strong>en</strong>tional system [13].<br />

Thermal Field-Flow Fractionation (Thermal FFF) is an excell<strong>en</strong>t technique for measuring<br />

Soret coeffici<strong>en</strong>ts particularly for dissolved polymers and susp<strong>en</strong><strong>de</strong>d particles [96].<br />

Flow<br />

Hot Wall<br />

Thermal<br />

diffusion<br />

Cold Wall<br />

13<br />

Flow<br />

Diffusion<br />

Fig. 1-4. Principle of Thermal Field-Flow Fractionation (ThFFF)<br />

1.4.4 Forced Rayleigh-Scattering Technique<br />

The principle of the forced Rayleigh scattering method is illustrated in Fig. 1-5. Two<br />

pulsed, high-power laser beams of equal wavel<strong>en</strong>gth and equal int<strong>en</strong>sity intersect in an<br />

absorbing sample. They g<strong>en</strong>erate an optical interfer<strong>en</strong>ce fringe pattern whose int<strong>en</strong>sity<br />

distribution is spatially sinusoidal. Following partial absorption of the laser light, this<br />

interfer<strong>en</strong>ce pattern induces a corresponding temperature grating, which in turn causes a<br />

conc<strong>en</strong>tration grating by the effect of thermal diffusion. Both gratings contribute to a<br />

combined refractive in<strong>de</strong>x grating that is read out by diffraction of a third laser beam.<br />

Analyzing the time <strong>de</strong>p<strong>en</strong>d<strong>en</strong>t diffraction effici<strong>en</strong>cy, three transport coeffici<strong>en</strong>ts can be<br />

obtained (the thermal diffusivity, the translation diffusion coeffici<strong>en</strong>t D, and the thermal<br />

diffusion coeffici<strong>en</strong>t DT). The ratio of the thermal diffusion coeffici<strong>en</strong>t and the translation<br />

diffusion coeffici<strong>en</strong>t allows the <strong>de</strong>termination of the Soret coeffici<strong>en</strong>t ST [113].


Fig. 1-5. Principle of forced Rayleigh scattering [99]<br />

1.4.5 The single-beam Z-scan or thermal l<strong>en</strong>s technique<br />

The z-scan is a simple technique for <strong>de</strong>termining the absorptive and refractive nonlinear<br />

optical properties of matter. In this type of technique a single laser beam is used for both<br />

heating and <strong>de</strong>tecting. Any effect that creates variation of the refractive in<strong>de</strong>x can be<br />

studied with this setup. Giglio and V<strong>en</strong>dramini, 1974 [36] noticed that, wh<strong>en</strong> an int<strong>en</strong>se<br />

narrow laser beam is reflected in a liquid, besi<strong>de</strong> the thermal expansion, the Soret effect<br />

appears. This work showed the effect of the laser beam in binary mixtures compared to<br />

pure liquids. This technique for <strong>de</strong>termination of the Soret coeffici<strong>en</strong>t is based on analysing<br />

the optical nonlinearities of the laser light.<br />

1.5 Conc<strong>en</strong>tration measurem<strong>en</strong>t<br />

A number of methods have be<strong>en</strong> used for measuring the change in composition resulting<br />

from thermal diffusion or diffusion. In some early investigations the gas was analysed by<br />

chemical methods, but for many mixtures there are more rapid and conv<strong>en</strong>i<strong>en</strong>t methods<br />

<strong>de</strong>p<strong>en</strong>ding on the variation with composition of properties such as thermal conductivity,<br />

viscosity and optical refractivity. The <strong>de</strong>velopm<strong>en</strong>t in rec<strong>en</strong>t years of Gas<br />

Chromatography-Mass Spectrometry (GC-MS) has <strong>en</strong>abled some progress to be ma<strong>de</strong>. In<br />

this section, we <strong>de</strong>scribe briefly these methods of measurem<strong>en</strong>t.<br />

14


1.5.1 From the variation of thermal conductivity<br />

An instrum<strong>en</strong>t was originally <strong>de</strong>vised by Shakespear in 1915 (see [27]) and as the<br />

instrum<strong>en</strong>t was primarily int<strong>en</strong><strong>de</strong>d to measure the purity of the air, the name<br />

"katharometer" was giv<strong>en</strong> to it. Katharometer [sometimes spelled “catherometer” and oft<strong>en</strong><br />

referred to as the thermal conductivity <strong>de</strong>tector (TCD) or the hot-wire <strong>de</strong>tector (HWD)]<br />

was applied by Ibbs (1921) in his first experim<strong>en</strong>ts on thermal diffusion.<br />

As we can see in Fig. 1-6, a typical kind of katharometer consists of a metal block in<br />

which one chamber is filled or purged with the gas mixture of unknown conc<strong>en</strong>tration and<br />

another one with a refer<strong>en</strong>ce gas. Each chamber contains a platinum filam<strong>en</strong>t forming a<br />

branch of a Wheatstone bridge circuit and heated by the bridge curr<strong>en</strong>t. The block serves as<br />

a heat sink at constant temperature. The katharometer conc<strong>en</strong>tration calibration is limited<br />

to a binary mixture. Therefore, this method is not appropriate in the case of more than two<br />

compon<strong>en</strong>ts.<br />

Heated<br />

metal block<br />

S<strong>en</strong>sor<br />

filam<strong>en</strong>t<br />

S<strong>en</strong>sor connections<br />

to Wheatstone<br />

bridge<br />

Refer<strong>en</strong>ce<br />

filam<strong>en</strong>t<br />

Gas from<br />

the bulb to<br />

measure Refer<strong>en</strong>ce<br />

gas<br />

Fig. 1-6. Diagram showing vertical section of the katharometer [27]<br />

15


Heat loss by radiation, convection, and leak through the supports is minimised in or<strong>de</strong>r to<br />

let the conduction through the gas be the dominant transfer mechanism of heat from the<br />

filam<strong>en</strong>t to the surroundings. Changes in gas composition in a chamber lead to temperature<br />

changes of the filam<strong>en</strong>t and thus to accompanying changes in resistance which are<br />

measured with the completed Wheatstone bridge. The heat lost from the filam<strong>en</strong>t will<br />

<strong>de</strong>p<strong>en</strong>d on both the thermal conductivity of the gas and its specific heat. Both these<br />

parameters will change in the pres<strong>en</strong>ce of a differ<strong>en</strong>t gas or solute vapor and as a result the<br />

temperature of the filam<strong>en</strong>t changes, causing a change in pot<strong>en</strong>tial across the filam<strong>en</strong>t. This<br />

pot<strong>en</strong>tial change is amplified and either fed to a suitable recor<strong>de</strong>r or passed to an<br />

appropriate data acquisition system. As the <strong>de</strong>tector filam<strong>en</strong>t is in thermal equilibrium with<br />

its surroundings and the <strong>de</strong>vice actually responds to the heat lost from the filam<strong>en</strong>t, the<br />

<strong>de</strong>tector is extremely flow and pressure s<strong>en</strong>sitive. Consequ<strong>en</strong>tly, all katharometer <strong>de</strong>tectors<br />

must be carefully thermostated and must be fitted with refer<strong>en</strong>ce cells to help comp<strong>en</strong>sate<br />

for changes in pressure or flow rate. Usually, one of the spirals of the katharometer is<br />

sealed perman<strong>en</strong>tly in air and the resistance readings are the refer<strong>en</strong>ce readings. Other<br />

filam<strong>en</strong>t is connected with the gas as analyze reading. The katharometer has the advantage<br />

that its op<strong>en</strong> cell can form part of the diffusion cell, and so it can indicate continuously the<br />

changes in composition as diffusion and thermal diffusion proceeds without sampling.<br />

1.5.2 From the variation of viscosity<br />

Van Itterbeek and van Paemel (1938, 1940) ([101, 102], see also [50]) have <strong>de</strong>veloped a<br />

method of measurem<strong>en</strong>t based on the damping of an oscillating-disk viscometers. The<br />

oscillating disk itself is a part of the top bulb of the thermal diffusion cells. The change in<br />

the composition of the upper part due to thermal diffusion, changes the viscosity of the<br />

mixture and th<strong>en</strong> corresponding change in composition is found from calibration curve.<br />

This method has a precision of the same as the conductivity <strong>de</strong>tector.<br />

1.5.3 Gas Chromatography (GC)<br />

In a Gas Chromatograph the sample is injected into a heated inlet where it is vaporized,<br />

and th<strong>en</strong> transferred into a chromatographic column. Differ<strong>en</strong>t compounds are separated in<br />

the column, primarily through their physical interaction with the walls of the column. Once<br />

16


separated, the compounds are fed into a <strong>de</strong>tector. There are two types of the <strong>de</strong>tector:<br />

Flame Ionization Detector (FID) and Electron Capture Detector (ECD). Schematic of a gas<br />

chromatograph flame ionization <strong>de</strong>tector is illustrated in Fig. 1-7. As we can see in this<br />

figure, GC-FID uses a flame ionization <strong>de</strong>tector for id<strong>en</strong>tification of compounds. The<br />

flame ionization <strong>de</strong>tector responds to compounds that create ions wh<strong>en</strong> combusted in a<br />

hydrog<strong>en</strong>-air flame. These ions pass into a <strong>de</strong>tector and are converted to an electrical<br />

signal. This method of analysis can be used for the <strong>de</strong>tection of compounds such as<br />

ethanol, acetal<strong>de</strong>hy<strong>de</strong>, ethyl acetate, and higher alcohols.<br />

Fig. 1-7. Schematics of a Gas Chromatograph Flame Ionization Detector (GC-FID)<br />

The ECD or electron capture <strong>de</strong>tector (Fig. 1-8) measures electron capturing compounds<br />

(usually halog<strong>en</strong>ated) by creating an electrical field in which molecules exiting a GC<br />

column can be <strong>de</strong>tected by the drop in curr<strong>en</strong>t in the field.<br />

Fig. 1-8. Schematics of a Gas Chromatograph Electron Capture Detector (GC-ECD)<br />

The ECD works by directing the gas phase output from the column across an electrical<br />

field applied across two electro<strong>de</strong>s, either using a constant DC pot<strong>en</strong>tial or a pulsed<br />

pot<strong>en</strong>tial. The electrical field is produced using a thermally stable 63Ni source that ionizes<br />

17


some of the carrier gas or auxiliary <strong>de</strong>tector gas (usually nitrog<strong>en</strong> or a mixture of argon<br />

95%, methane 5%) and produces a curr<strong>en</strong>t betwe<strong>en</strong> a biased pair of electro<strong>de</strong>s. The ECD<br />

is one of the most s<strong>en</strong>sitive gas chromatography <strong>de</strong>tectors available. The s<strong>en</strong>sitivity of the<br />

ECD <strong>en</strong>ables it to provi<strong>de</strong> unmatched performance for extremely tough applications. It is<br />

the first choice for certain <strong>en</strong>vironm<strong>en</strong>tal chromatography applications due to its extreme<br />

s<strong>en</strong>sitivity to halog<strong>en</strong>ated compounds like PCBs (Polychlorinated biph<strong>en</strong>yls),<br />

organochlorine pestici<strong>de</strong>s, herbici<strong>de</strong>s, and halog<strong>en</strong>ated hydrocarbons. The ECD is 10-1000<br />

time more s<strong>en</strong>sitive than the FID (Flame Ionization Detector), but has a limited dynamic<br />

range and finds its greatest application in analysis of halog<strong>en</strong>ated compounds.<br />

1.5.4 Analysis by mass spectrometer<br />

Mass spectrometers are s<strong>en</strong>sitive <strong>de</strong>tectors of isotopes based on their masses. For the study<br />

of thermal diffusion in isotopic mixtures, a mass spectrometer is necessary. In the mass<br />

spectrometer the mixture is first ionized by passage through an electron beam as shown in<br />

Fig. 1-9; the ions are accelerated by an electric field and th<strong>en</strong> passed through a slit system<br />

into a magnetic field by which they are <strong>de</strong>flected through an angle which <strong>de</strong>p<strong>en</strong>ds on the<br />

mass and velocity. The final elem<strong>en</strong>t of the mass spectrometer is the <strong>de</strong>tector. The <strong>de</strong>tector<br />

records either the charge induced or the curr<strong>en</strong>t produced wh<strong>en</strong> an ion passes by or hits a<br />

surface. The combination of a mass spectrometer and a gas chromatograph makes a<br />

powerful tool for the <strong>de</strong>tection of trace quantities of contaminants or toxins.<br />

Fig. 1-9. Schematics of a simple mass spectrometer<br />

18


1.6 Conclusion<br />

From the discussion in section 1.3, it is clear that mo<strong>de</strong>ls of transport in porous media are<br />

related to scale <strong>de</strong>scription. The prediction and mo<strong>de</strong>lling of fluid flow processes in the<br />

subsurface is necessary e.g. for groundwater remediation or oil recovery. In most<br />

applications the fluid flow is <strong>de</strong>termined by, in g<strong>en</strong>eral, highly heterog<strong>en</strong>eous distribution<br />

of the soil properties. The conclusion of the <strong>de</strong>tailed knowledge of the heterog<strong>en</strong>eous<br />

parameter distribution into a flow mo<strong>de</strong>l is computationally not feasible. It is therefore an<br />

important task to <strong>de</strong>velop upscaling methods to simplify the small-scale flow mo<strong>de</strong>l while<br />

still including the impact of the heterog<strong>en</strong>eities as far as possible. In this study we have<br />

used the volume averaging technique to obtain the macro-scale properties of the porous<br />

media because this method has be<strong>en</strong> proved to be suitable tool for mo<strong>de</strong>lling transport<br />

ph<strong>en</strong>om<strong>en</strong>a in heterog<strong>en</strong>eous porous media.<br />

Differ<strong>en</strong>t experim<strong>en</strong>tal techniques, which permit to measure the separation and therefore<br />

calculate the thermal diffusion coeffici<strong>en</strong>ts, have be<strong>en</strong> pres<strong>en</strong>ted. It follows that, two-bulb<br />

method, among other methods, is a suitable method for this study since we can measure the<br />

both diffusion and thermal diffusion coeffici<strong>en</strong>ts. It is easily adoptable to apply for a<br />

porous medium case. In this method thermal diffusion process does not disturb by the free<br />

convection which is negligible in this system.<br />

Katharometer, <strong>de</strong>spite its limitation to binary mixtures, is still most commonly used<br />

<strong>de</strong>tector in many industries. Katharometer is simple in <strong>de</strong>sign and requires minimal<br />

electronic support and, as a consequ<strong>en</strong>ce, is also relatively inexp<strong>en</strong>sive compared with<br />

other <strong>de</strong>tectors. Its op<strong>en</strong> cell can form part of the diffusion cell, and so it can indicate<br />

continuously the changes in composition without sampling. This is why that in this study,<br />

we have used a conductivity <strong>de</strong>tector method with katharometer to analyze the separation<br />

process in a two-bulb method.<br />

19


Chapter 2<br />

Theoretical Predictions of the Effective Diffusion<br />

and Thermal diffusion Coeffici<strong>en</strong>ts<br />

in Porous Media


2. Theoretical predictions of the effective diffusion and thermal<br />

diffusion coeffici<strong>en</strong>ts in porous media<br />

This chapter pres<strong>en</strong>ts the theoretical <strong>de</strong>termination of the effective Darcy-scale coeffici<strong>en</strong>ts<br />

for heat and mass transfer in porous media, including the thermal diffusion effect, using a<br />

volume averaging technique. The closure problems related to the pore-scale physics are<br />

solved over periodic unit cells repres<strong>en</strong>tative of the porous structure.<br />

Nom<strong>en</strong>clature of Chapter 2<br />

a v<br />

A βσ V , interfacial area per unit<br />

volume, m -1<br />

A 0<br />

Specific surface area, m -1 p β<br />

β<br />

A βσ<br />

Area of the β-σ interface contained<br />

within the macroscopic region, m 2<br />

r<br />

A βe<br />

Area of the <strong>en</strong>trances and exits of the βσ<br />

phase associated with the<br />

macroscopic system, m 2<br />

A Area of the β-σ interface within the<br />

βσ<br />

averaging volume, m 2<br />

Gas slip factor<br />

b i<br />

b Cβ<br />

Mapping vector field for β<br />

21<br />

p β<br />

r β<br />

Sc<br />

S T<br />

c ~ , m s β<br />

b Sβ<br />

Mapping vector field for c β<br />

~ , m s σ<br />

b x-coordinate coeffici<strong>en</strong>t of b Sβ<br />

b<br />

Sβ<br />

x<br />

Sββ<br />

Vector field that maps ∇ onto<br />

c β<br />

~ ,m<br />

b Sβσ<br />

Vector field that maps<br />

b Tβ<br />

bTβ<br />

x<br />

c β<br />

~ ,m<br />

Tβ<br />

β<br />

σ<br />

∇ onto<br />

Tσ<br />

Mapping vector field for T β<br />

~ , m<br />

*<br />

ST *<br />

ST<br />

xx<br />

t Time, s<br />

*<br />

t<br />

x-coordinate coeffici<strong>en</strong>t of b Tβ<br />

T β<br />

Pressure in the β-phase, Pa<br />

Intrinsic average pressure in the βphase,<br />

Pa<br />

Position vector, m<br />

Scalar field that maps<br />

⎜<br />

⎛ T β<br />

⎝<br />

β<br />

−<br />

σ<br />

T ⎟<br />

⎞<br />

σ onto c β<br />

⎠<br />

~<br />

Schmidt number<br />

Soret number, 1/K<br />

Scalar field that maps<br />

⎜<br />

⎛ T β<br />

⎝<br />

β<br />

−<br />

σ<br />

T ⎟<br />

⎞<br />

σ onto T β<br />

⎠<br />

~<br />

Scalar field that maps<br />

⎜<br />

⎛ T β<br />

⎝<br />

β<br />

−<br />

σ<br />

T ⎟<br />

⎞<br />

σ onto T σ<br />

⎠<br />

~<br />

Effective Soret number, 1/K<br />

Longitudinal Soret number, 1/K<br />

Characteristic process time, s<br />

Temperature of the β-phase, K


Tββ<br />

Vector field that maps ∇ onto<br />

T β<br />

~ ,m<br />

b Tβσ<br />

Vector field that maps<br />

bTσβ<br />

bTσσ<br />

c p<br />

c β<br />

c β<br />

β<br />

T β<br />

~ ,m<br />

Tβ<br />

Tσ<br />

β<br />

σ<br />

∇ onto<br />

Vector field that maps ∇ onto<br />

T σ<br />

~ ,m<br />

Tβ<br />

Vector field that maps ∇ onto<br />

T σ<br />

~ ,m<br />

Tσ<br />

Constant pressure heat capacity, J.kg/K<br />

Total mass fraction in the β-phase<br />

Intrinsic average mass fraction in the βphase<br />

c β<br />

~ Spatial <strong>de</strong>viation mass fraction in the βphase<br />

Binary diffusion coeffici<strong>en</strong>t, m 2 /s<br />

D β<br />

D Thermal diffusion coeffici<strong>en</strong>t, m<br />

Tβ<br />

2 /s.K<br />

*<br />

D Total thermal diffusion t<strong>en</strong>sor, m<br />

T β<br />

2 /s.K<br />

*<br />

D Tβ<br />

*<br />

D ββ<br />

xx<br />

Longitudinal thermal diffusion<br />

coeffici<strong>en</strong>t, m 2 /s.K<br />

Effective thermal diffusion t<strong>en</strong>sor<br />

T β<br />

associated with ∇ T in the β-phase<br />

*<br />

D βσ<br />

Effective thermal diffusion t<strong>en</strong>sor<br />

T σ<br />

associated with ∇ T in the β-phase<br />

*<br />

D β<br />

*<br />

D β<br />

F<br />

xx<br />

Total dispersion t<strong>en</strong>sor, m 2 /s<br />

Longitudinal dispersion coeffici<strong>en</strong>t,<br />

m 2 /s<br />

β<br />

σ<br />

β<br />

σ<br />

22<br />

T β<br />

T β<br />

~<br />

u Cβ<br />

β<br />

Intrinsic average temperature in the<br />

β-phase, K<br />

Spatial <strong>de</strong>viation temperature , K<br />

One-equation mo<strong>de</strong>l mass transport<br />

coeffici<strong>en</strong>t associated with<br />

β<br />

⎜<br />

⎛ σ<br />

∇.<br />

T − ⎟<br />

⎞<br />

β Tσ<br />

in the β-phase<br />

⎝<br />

⎠<br />

equation<br />

u Two-equation mo<strong>de</strong>l heat transport<br />

ββ<br />

β<br />

coeffici<strong>en</strong>t associated with ∇ Tβ<br />

in<br />

the β-phase equation<br />

u βσ<br />

Two-equation mo<strong>de</strong>l heat transport<br />

σ<br />

coeffici<strong>en</strong>t associated with ∇ Tσ<br />

in<br />

the β-phase equation<br />

u Two-equation mo<strong>de</strong>l heat transport<br />

σβ<br />

β<br />

coeffici<strong>en</strong>t associated with ∇ Tβ<br />

in<br />

the σ-phase equation<br />

u Two-equation mo<strong>de</strong>l heat transport<br />

σσ<br />

σ<br />

coeffici<strong>en</strong>t associated with ∇ Tσ<br />

in<br />

the σ-phase equation<br />

P<br />

Mean pressure, Pa<br />

p Cβ<br />

Capillary pressure, Pa<br />

Pe Cell Péclet number<br />

Pr<br />

r 0<br />

v β<br />

v β<br />

v~<br />

β<br />

V β<br />

β<br />

Prandtl number<br />

Radius of the averaging volume, m<br />

Mass average velocity in the β-phase,<br />

m/s<br />

Intrinsic average mass average<br />

velocity in the β-phase, m/s<br />

Spatial <strong>de</strong>viation mass average<br />

velocity, m/s<br />

Volume of the β-phase contained<br />

within the averaging volume, m 3<br />

Forchheimer correction t<strong>en</strong>sor V Local averaging volume, m 3


g Gravitational acceleration, m 2 /s y<br />

h<br />

I<br />

Film heat transfer coeffici<strong>en</strong>t, J<br />

Unit t<strong>en</strong>sor<br />

m.<br />

s.<br />

K z<br />

Greek symbols<br />

k a<br />

Appar<strong>en</strong>t gas permeability, m 2<br />

k Relative permeability<br />

rβ<br />

k β<br />

k σ<br />

K β<br />

Thermal conductivity of the fluid phase,<br />

W/m.K<br />

Thermal conductivity of the solid phase,<br />

W/m.K<br />

Permeability t<strong>en</strong>sor, m 2<br />

k Two-equation mo<strong>de</strong>l effective thermal<br />

ββ<br />

conductivity t<strong>en</strong>sor associated with<br />

β<br />

∇ in the β-phase equation<br />

Tβ<br />

k Two-equation mo<strong>de</strong>l effective thermal<br />

βσ<br />

conductivity t<strong>en</strong>sor associated with<br />

σ<br />

∇ in the σ-phase equation<br />

Tσ<br />

k Two-equation mo<strong>de</strong>l effective thermal<br />

σβ<br />

conductivity t<strong>en</strong>sor associated with<br />

β<br />

∇ in the σ-phase equation<br />

Tβ<br />

k Two-equation mo<strong>de</strong>l effective thermal<br />

σσ<br />

conductivity t<strong>en</strong>sor associated with<br />

k , k<br />

*<br />

β<br />

*<br />

σ<br />

∇ in the σ-phase equation<br />

Tσ<br />

Total thermal conductivity t<strong>en</strong>sors for<br />

no-conductive and conductive solid<br />

phase, W/m.K<br />

23<br />

x, Cartesian coordinates, m<br />

Elevation in the gravitational field, m<br />

β F<br />

ε β<br />

κ σ kβ<br />

A factor experim<strong>en</strong>tally <strong>de</strong>duced<br />

Volume fraction of the β-phase or<br />

porosity<br />

k , conductivity ratio<br />

λ Mean free path of gas, μm<br />

μ β<br />

β<br />

Dynamic viscosity for the β-phase,<br />

Pa.s<br />

μ ~ Effective viscosity, Pas.s<br />

υ β<br />

ρ β<br />

τ<br />

Kinematic viscosity for the β-phase,<br />

m 2 /s<br />

Total mass d<strong>en</strong>sity in the β-phase,<br />

kg/m 3<br />

Scalar tortuosity factor<br />

ϕ Arbitrary function<br />

*<br />

k β<br />

xx<br />

Longitudinal thermal dispersion<br />

coeffici<strong>en</strong>t, W/m.K<br />

ψ Separation factor or dim<strong>en</strong>sionless<br />

Soret number<br />

k Klink<strong>en</strong>berg permeability, W/m.K Subscripts, superscripts and other symbols<br />

∞<br />

*<br />

k ∞<br />

l<br />

Asymptotic thermal dispersion<br />

coeffici<strong>en</strong>t , W/m.K<br />

Characteristic l<strong>en</strong>gth associated with the<br />

microscopic scale, m<br />

l Characteristic l<strong>en</strong>gth scale associated<br />

UC<br />

with a unit cell, m<br />

l β<br />

L<br />

Characteristic l<strong>en</strong>gth for the β-phase, m<br />

Characteristic l<strong>en</strong>gth for macroscopic<br />

quantities, m<br />

L Characteristic l<strong>en</strong>gth for∇ ε , m *<br />

ε<br />

ref<br />

β<br />

Refers to the refer<strong>en</strong>ce gas<br />

Fluid-phase<br />

σ Solid-phase<br />

βσ<br />

βe<br />

β-σ interphase<br />

Fluid-phase <strong>en</strong>trances and exits<br />

Effective quantity


M<br />

Gas molecular weight, g/mol<br />

n Unit normal vector directed from the β-<br />

βσ<br />

phase toward the σ –phase<br />

24<br />

β<br />

Spatial average<br />

Intrinsic β-phase average


2.1 Introduction<br />

It is well established, see for instance [39], that a multicompon<strong>en</strong>t system un<strong>de</strong>r<br />

nonisothermal condition is subject to mass transfer related to coupled-transport<br />

ph<strong>en</strong>om<strong>en</strong>a. This has strong practical importance in many situations since the flow<br />

dynamics and convective patterns in mixtures are more complex than those of onecompon<strong>en</strong>t<br />

fluids due to the interplay betwe<strong>en</strong> advection and mixing, solute diffusion, and<br />

the Soret effect (or thermal diffusion) [112]. The Soret coeffici<strong>en</strong>t may be positive or<br />

negative <strong>de</strong>p<strong>en</strong>ding on the direction of migration of the refer<strong>en</strong>ce compon<strong>en</strong>t (to the cold<br />

or to the hot region).<br />

There are many important processes in nature and technology where thermal diffusion<br />

plays a crucial role. Thermal diffusion has various technical applications, such as isotope<br />

separation in liquid and gaseous mixtures [86, 87], polymer solutions and colloidal<br />

dispersions [112], study of compositional variation in hydrocarbon reservoirs [32], coating<br />

of metallic items, etc. It also affects compon<strong>en</strong>t separation in oil wells, solidifying metallic<br />

alloys, volcanic lava, and in the Earth Mantle [45].<br />

Platt<strong>en</strong> and Costesèque (2004) searched the response to the basic question:” is the Soret<br />

coeffici<strong>en</strong>t the same in a free fluid and in a porous medium?” They measured separately<br />

four coeffici<strong>en</strong>ts: isothermal diffusion and thermal diffusion coeffici<strong>en</strong>ts, both in free fluid<br />

and porous media. They measured the diffusion coeffici<strong>en</strong>t in free fluid by the op<strong>en</strong>-<strong>en</strong><strong>de</strong>dcapillary<br />

(OEC) technique, and th<strong>en</strong> they g<strong>en</strong>eralized the same OEC technique to porous<br />

media. The thermal diffusion coeffici<strong>en</strong>t in the free system has also be<strong>en</strong> measured by the<br />

thermogravitational column technique [73]. The thermal diffusion coeffici<strong>en</strong>t of the same<br />

mixture was <strong>de</strong>termined in a porous medium by the same technique, except that they filled<br />

the gap betwe<strong>en</strong> two conc<strong>en</strong>tric cylin<strong>de</strong>rs with zirconia spheres. In spite of the small errors<br />

that they had on the Soret coeffici<strong>en</strong>t due to measuring in<strong>de</strong>p<strong>en</strong>d<strong>en</strong>tly diffusion and<br />

thermal diffusion coeffici<strong>en</strong>t they announced that the Soret coeffici<strong>en</strong>t is the same in a free<br />

fluid and in porous medium [74]. The experim<strong>en</strong>tal study of Costesèque et al. (2004) for a<br />

horizontal Soret-type thermal diffusion cell, filled first with the free liquid and next with a<br />

porous medium showed also that the results are not significantly differ<strong>en</strong>t [20].<br />

Saghir et al. (2005) have reviewed some aspects of thermal diffusion in porous media;<br />

including the theory and the numerical procedure which have be<strong>en</strong> <strong>de</strong>veloped to simulate<br />

these ph<strong>en</strong>om<strong>en</strong>a [91]. In many other works on thermal diffusion in a square porous cavity,<br />

25


the thermal diffusion coeffici<strong>en</strong>t in free fluid almost has be<strong>en</strong> used instead of an effective<br />

coeffici<strong>en</strong>t containing the tortuosity and dispersion effect. Therefore, there are many<br />

discrepancies betwe<strong>en</strong> the predictions and measurem<strong>en</strong>ts separation.<br />

The effect of dispersion on effective diffusion is now well established (see for example<br />

Saffman (1959), Bear (1972), …) but this effect on thermal diffusion has received limited<br />

att<strong>en</strong>tion. Fargue et al. (1998) searched the <strong>de</strong>p<strong>en</strong>d<strong>en</strong>ce of the effective thermal diffusion<br />

coeffici<strong>en</strong>t on flow velocity in a porous packed thermogravitational column. They showed<br />

that the effective thermal diffusion coeffici<strong>en</strong>t in thermogravitational column filled with<br />

porous media inclu<strong>de</strong>s a <strong>de</strong>p<strong>en</strong>d<strong>en</strong>cy upon the fluid velocity. Their results showed that the<br />

behaviour of the effective thermal diffusion coeffici<strong>en</strong>t looks very similar to the effective<br />

diffusion coeffici<strong>en</strong>t in porous media [31].<br />

The numerical mo<strong>de</strong>l of Nasrabadi and Firoozabadi (2006) in a packed thermogravitational<br />

column was not able to reveal a dispersion effect on the thermal diffusion process, perhaps<br />

mainly due to low velocities [66].<br />

In this chapter, we have used the volume averaging method which has be<strong>en</strong> ext<strong>en</strong>sively<br />

used to predict the effective isothermal transport properties in porous media. The<br />

consi<strong>de</strong>red media can also be subjected to thermal gradi<strong>en</strong>ts coming from natural origin<br />

(geothermal gradi<strong>en</strong>ts, intrusions,…) or from anthropic anomalies (waste storages,…).<br />

Thermal diffusion has rarely be<strong>en</strong> tak<strong>en</strong> completely un<strong>de</strong>r consi<strong>de</strong>ration, in most<br />

<strong>de</strong>scription, coupled effects being g<strong>en</strong>erally forgott<strong>en</strong> or neglected. However, the pres<strong>en</strong>ce<br />

of temperature gradi<strong>en</strong>t in the medium can g<strong>en</strong>erate a mass flux.<br />

For mo<strong>de</strong>lling mass transfer by thermal diffusion, the effective thermal conductivity must<br />

be first <strong>de</strong>termined. Differ<strong>en</strong>t mo<strong>de</strong>ls have be<strong>en</strong> investigated for two-phase heat transfer<br />

systems <strong>de</strong>p<strong>en</strong>ding on the validity of the local thermal equilibrium assumption. Wh<strong>en</strong> one<br />

accepts this assumption, macroscopic heat transfer can be <strong>de</strong>scribed correctly by a classical<br />

one-equation mo<strong>de</strong>l [47, 82, 79, 84]. The rea<strong>de</strong>r can look at [3] for the possible impact of<br />

non-equilibrium on various flow conditions. For many initial boundary-value problems, the<br />

two-equation mo<strong>de</strong>l shows an asymptotic behaviour that can be mo<strong>de</strong>lled with a “non<br />

equilibrium” one-equation mo<strong>de</strong>l [115, 116]. The resulting thermal dispersion t<strong>en</strong>sor is<br />

greater than the one-equation local-equilibrium dispersion t<strong>en</strong>sor. It can also be obtained<br />

through a special closure problem as shown in [64]. These mo<strong>de</strong>ls can also be ext<strong>en</strong><strong>de</strong>d to<br />

more complex situations like two-phase flow [65], reactive transport [77, 85, 93].<br />

26


However, for all these mo<strong>de</strong>ls many coupling ph<strong>en</strong>om<strong>en</strong>a have be<strong>en</strong> discar<strong>de</strong>d in the<br />

upscaling analysis. This is particularly the case for the possible coupling with the transport<br />

of constitu<strong>en</strong>ts in the case of mixtures.<br />

A mo<strong>de</strong>l for Soret effect in porous media has be<strong>en</strong> <strong>de</strong>termined by Lacabanne et al (2002),<br />

they pres<strong>en</strong>ted a homog<strong>en</strong>ization technique for <strong>de</strong>termining the macroscopic Soret number<br />

in porous media. They assumed a periodic porous medium with the periodical repetition of<br />

an elem<strong>en</strong>tary cell. In this mo<strong>de</strong>l, the effective thermal diffusion and isothermal diffusion<br />

coeffici<strong>en</strong>t are calculated by only one closure problem while, in this study, two closure<br />

problems have to be solved separately to obtain effective isothermal and thermal diffusion<br />

coeffici<strong>en</strong>ts. They have also studied the local coupling betwe<strong>en</strong> velocity and Soret effect in<br />

a tube with a thermal gradi<strong>en</strong>t. The results of this mo<strong>de</strong>l showed that wh<strong>en</strong> convection is<br />

coupled with Soret effect, diffusion removes the negative part of the separation profile<br />

[51]. However, they calculated the effective coeffici<strong>en</strong>ts for a purely diffusive regime for<br />

which one cannot observe the effect of force convection and conductivity ratio as<br />

explained later in this study. In addition, these results have not be<strong>en</strong> validated with<br />

experim<strong>en</strong>tal results or a direct pore-scale numerical approach.<br />

In this chapter, effective properties will be calculated for a simple unit-cell but for various<br />

physical parameters, in particular the Péclet number and the thermal conductivity ratio.<br />

2.2 Governing microscopic equation<br />

We consi<strong>de</strong>r in this study a binary mixture fluid flowing through a porous medium<br />

subjected to a thermal gradi<strong>en</strong>t. This system is illustrated in Fig. 2-1, the fluid phase is<br />

id<strong>en</strong>tified as the β-phase while the rigid and impermeable solid is repres<strong>en</strong>ted by the σ-<br />

phase.<br />

From the thermodynamics of irreversible processes as originally formulated by Onsager<br />

(1931) the diagonal effects that <strong>de</strong>scribe heat and mass transfer are Fourier’s law which<br />

relates heat flow to the temperature gradi<strong>en</strong>t and Fick’s law which relates mass flow to the<br />

conc<strong>en</strong>tration gradi<strong>en</strong>t. There are also cross effects or coupled-processes: the Dufour effect<br />

quantifies the heat flux caused by the conc<strong>en</strong>tration gradi<strong>en</strong>t and the Soret effect, the mass<br />

flux caused by the temperature gradi<strong>en</strong>t.<br />

27


Fig. 2-1. Problem configuration<br />

In this study, we neglect the Dufour effect, which is justified in liquids [75] but in gaseous<br />

mixtures the Dufour coupling may becomes more and more important and can change the<br />

stability behaviour of the mixture in a Rayleigh-Bénard problem in comparison to liquid<br />

mixtures [43].<br />

Therefore, the transport of <strong>en</strong>ergy at the pore level is <strong>de</strong>scribed by the following equations<br />

and boundary conditions for the fluid (β-phase) and solid (σ-phase)<br />

∂T<br />

β ( ρ ) + ( ρc<br />

) ∇ ( T ) = ∇.<br />

( k ∇T<br />

)<br />

cp β<br />

p<br />

∂t<br />

β<br />

. v , in the β-phase ( 2-1)<br />

β<br />

β<br />

β<br />

β<br />

BC1: T β = Tσ<br />

, at A βσ<br />

( 2-2)<br />

BC2: . ( k ∇ T ) = n . ( k ∇T<br />

)<br />

n βσ β β βσ σ σ , at βσ<br />

∂Tσ<br />

( ρ ) = ∇ ( k ∇T<br />

)<br />

cp σ<br />

∂t<br />

σ<br />

σ<br />

A ( 2-3)<br />

. , in the σ-phase ( 2-4)<br />

where A βσ is the area of the β-σ interface contained within the macroscopic region.<br />

We assume in this work that the physical properties of the fluid and solid are constant.<br />

Th<strong>en</strong> the compon<strong>en</strong>t pore-scale mass conservation is <strong>de</strong>scribed by the following equation<br />

and boundary conditions for the fluid phase [9]<br />

∂c<br />

∂t<br />

β<br />

( c ) = ∇.<br />

( D ∇c<br />

+ D T )<br />

+ ∇.<br />

v , in the β-phase ( 2-5)<br />

β<br />

β<br />

β<br />

β<br />

Tβ ∇<br />

β<br />

28


At the fluid-solid interfaces there is no transport of solute so that the mass flux (the sum of<br />

diffusion and thermal diffusion flux) is zero<br />

BC1: . ( c + D ∇T<br />

) = 0<br />

n βσ Dβ ∇ β Tβ<br />

β , at A βσ<br />

( 2-6)<br />

where c β is the mass fraction of one compon<strong>en</strong>t in the β-phase, Dβ and D Tβ<br />

are the<br />

molecular isothermal diffusion coeffici<strong>en</strong>t and thermal diffusion coeffici<strong>en</strong>t. n βσ is unit<br />

normal from the liquid to the solid phase. We neglect any accumulation and reaction of<br />

solute at the fluid-solid interface as well as the ph<strong>en</strong>om<strong>en</strong>on of surface diffusion.<br />

To <strong>de</strong>scribe completely the problem, the equations of continuity and motion have to be<br />

introduced for the fluid phase. We use Stokes equation for the flow motion at the porescale,<br />

assuming classically negligible inertia effects, also named creeping flow. This is a<br />

type of fluid flow where advective inertial forces are small compared to viscous forces (the<br />

Reynolds number is low, i.e. Re


ϕ<br />

β<br />

=<br />

1<br />

V<br />

∫<br />

Vβ<br />

ϕ dV<br />

β<br />

while the second average is the intrinsic average <strong>de</strong>fined by<br />

ϕ<br />

β<br />

β<br />

=<br />

1<br />

∫<br />

Vβ V<br />

ϕ dV<br />

β<br />

β<br />

30<br />

( 2-10)<br />

( 2-11)<br />

Here we have used V β to repres<strong>en</strong>t the volume of the β-phase contained within the<br />

averaging volume. These two averages are related by<br />

β<br />

β<br />

β<br />

β<br />

ϕ = ε ϕ<br />

( 2-12)<br />

in which ε β is the volume fraction of the β-phase or porosity in the one phase flow case.<br />

The phase or superficial averages are volume fraction <strong>de</strong>p<strong>en</strong>d<strong>en</strong>t. From the diagram in Fig.<br />

2-1 we can see that the sum of volume fractions of the two phases satisfies<br />

ε ε = 1<br />

( 2-13)<br />

β + σ<br />

In or<strong>de</strong>r to carry out the necessary averaging procedures to <strong>de</strong>rive governing differ<strong>en</strong>tial<br />

equations for the intrinsic average fields, we need to make use of the spatial averaging<br />

theorem, writt<strong>en</strong> here for any g<strong>en</strong>eral scalar quantity ϕ β associated with the β-phase<br />

1<br />

∇ϕ<br />

β = ∇ ϕβ<br />

+ ∫ n<br />

V<br />

A<br />

βσ<br />

βσ<br />

ϕ dA<br />

β<br />

( 2-14)<br />

A similar equation may be writt<strong>en</strong> for any fluid property associated with the β-phase. Note<br />

that the area integral in equation ( 2-14) involves unit normal from the β-phase to the σ –<br />

phase. In writing corresponding equation for the σ –phase, we realize that n βσ = −nσβ<br />

according to the <strong>de</strong>finitions of the unit normal. Following classical i<strong>de</strong>as [111] we will try<br />

to solve approximately the problems in terms of averaged values and <strong>de</strong>viations.<br />

The pore-scale fields <strong>de</strong>viation in the β-phase and σ -phase are respectively <strong>de</strong>fined by<br />

ϕ<br />

~<br />

β<br />

σ<br />

β = ϕ β + ϕ β and ϕ σ ϕσ<br />

+ ϕσ<br />

=<br />

~<br />

( 2-15)<br />

The classical l<strong>en</strong>gth-scale constraints (Fig. 2-1) have be<strong>en</strong> imposed by assuming<br />

l β


2.4 Darcy’s law<br />

If we assume that d<strong>en</strong>sity and viscosity are constants, the flow problem can be solved<br />

in<strong>de</strong>p<strong>en</strong>d<strong>en</strong>tly from the heat and constitu<strong>en</strong>t transport equations, and the change of scale<br />

for Stokes flow equation and continuity has already be<strong>en</strong> investigated and this leads to<br />

Darcy’s law and the volume averaged continuity equation [94, 110] writt<strong>en</strong> as<br />

β ( ∇ β − ρ g)<br />

K β<br />

v β = − . p<br />

β , in the porous medium ( 2-17)<br />

μ<br />

β<br />

∇. v β = 0 , in the porous medium ( 2-18)<br />

where K β is the permeability t<strong>en</strong>sor.<br />

Note that the Darcy velocity, v β , is a superficial velocity based on the <strong>en</strong>tire volume, not<br />

just the fluid volume. One can also related the Darcy velocity to the average intrinsic<br />

velocity,<br />

β<br />

β<br />

v as<br />

β<br />

β ε β vβ<br />

v = ( 2-19)<br />

Values of the liquid-phase permeability vary wi<strong>de</strong>ly, from<br />

down to<br />

18<br />

10 −<br />

to<br />

10<br />

−20 2<br />

31<br />

7<br />

10 − to<br />

10<br />

m for clean gravel<br />

−9 2<br />

m for granite (Bear, 1979 [7]). A Darcy (or Darcy unit) and<br />

millidarcies (mD) are units of permeability, named after H<strong>en</strong>ry Darcy. These units are<br />

wi<strong>de</strong>ly used in petroleum <strong>en</strong>gineering and geology. The unit Darcy is equal to<br />

0. 987 m<br />

−12<br />

2<br />

× 10 but most of the times it is simply assumed<br />

1D m<br />

−12<br />

2<br />

= 10 .<br />

Darcy’s law is applicable to low velocity flow, which is g<strong>en</strong>erally the case in porous media<br />

flow, and to regions without boundary shear flow, such as away from walls. Wh<strong>en</strong> wall<br />

shear is important, the Brinkman ext<strong>en</strong>sion can be used as discussed below. A Forchheimer<br />

equation is appropriate wh<strong>en</strong> the inertial effect is important. In some situations (e.g., Vafai<br />

and Ti<strong>en</strong>, 1981), the Brinkman and Forchheimer equations are both employed. One must<br />

use an effective correlation of appar<strong>en</strong>t gas permeability in tight porous media because of<br />

Knuds<strong>en</strong> effect.<br />

2.4.1 Brinkman term<br />

The Brinkman ext<strong>en</strong>sion to Darcy’s law equation (introduced by Brinkman in 1947)<br />

inclu<strong>de</strong>s the effect of wall or boundary shear on the flow velocity, or


β μ β<br />

2<br />

0 = −∇ p β + ρ β g − v ~<br />

β + μ β ∇ v β<br />

( 2-20)<br />

K<br />

β<br />

the third term on the RHS is a shear stress term such as would be required by no-slip<br />

condition. The coeffici<strong>en</strong>t μ β<br />

~ is an effective viscosity, which in g<strong>en</strong>eral is not equal to the<br />

fluid viscosity, μ β , as discussed by Nield and Bejan (1999) [70]. For many situations, the<br />

use of the boundary shear term is not necessary. Without discussing the validity of<br />

Brinkman’s equation near a wall or in areas of rapid porosity variations, the effect is only<br />

significant in a region close to the boundary whose thickness is of or<strong>de</strong>r of the square root<br />

0.<br />

5<br />

of the gas permeability, K β , (assuming ~ μ β = μβ<br />

), so for most applications and also in<br />

this study the effect can be ignored.<br />

The Brinkman equation is also oft<strong>en</strong> employed at the interface betwe<strong>en</strong> a porous medium<br />

and a free fluid (fluid with no porous medium), in or<strong>de</strong>r to obtain continuity of shear stress<br />

(more <strong>de</strong>tail in [70] and [47])<br />

2.4.2 No-linear case<br />

At low pore velocities, Darcy’s law works quite well. However, as the pore velocities<br />

increase, the inertial effect becomes very important, the flow resistance becomes nonlinear,<br />

and the Forchheimer equation is more appropriate as<br />

β μ<br />

0 = −∇ p + ρ − v v<br />

( 2-21)<br />

β<br />

β<br />

β g − v β ρ β β F<br />

K β<br />

β<br />

The third term on the RHS is a nonlinear flow resistance term. According to Nield and<br />

Bejan (1999), the above equation is based on the work of Dupuit (1863) and Forchheimer<br />

(1901) as modified by Ward (1964). β F is a factor to be experim<strong>en</strong>tally <strong>de</strong>duced.<br />

Whitaker, (1996) <strong>de</strong>rived Darcy's law with the Forchheimer correction for homog<strong>en</strong>eous<br />

porous media using the method of volume averaging. Beginning with the Navier-Stokes<br />

equations, they found that the volume averaged mom<strong>en</strong>tum equation to be giv<strong>en</strong> by<br />

β ( ∇ β − ρ β g)<br />

F v β<br />

K<br />

= p ( 2-22)<br />

β<br />

v β − . − .<br />

μβ<br />

32<br />

β


where F is the Forchheimer correction t<strong>en</strong>sor. In this equation, K β and F are <strong>de</strong>termined<br />

by closure problems that must be solved using a spatially periodic mo<strong>de</strong>l of a porous<br />

medium [110].<br />

2.4.3 Low permeability correction<br />

Based on Darcy’s law, the mass flux for a giv<strong>en</strong> pressure drop should <strong>de</strong>crease as the<br />

average pressure is reduced due to the change in gas d<strong>en</strong>sity. However, Knuds<strong>en</strong> found that<br />

at low pressures, the mass flux reaches a minimum value and th<strong>en</strong> increases with<br />

<strong>de</strong>creasing pressure, which is due to slip, or the fact that the fluid velocity at the wall is not<br />

zero due to free-molecule flow. As the capillary tubes get smaller and smaller, the gas<br />

molecular mean free path becomes of the same or<strong>de</strong>r, and free molecule, or Knuds<strong>en</strong>,<br />

diffusion becomes important.<br />

Assuming gas flow in an i<strong>de</strong>alized porous medium, using Poiseuille's law or Darcy's law, a<br />

correlation betwe<strong>en</strong> the appar<strong>en</strong>t and “true” permeability of a porous medium was <strong>de</strong>rived<br />

as (Klink<strong>en</strong>berg, 1941)<br />

k a = k∞<br />

⎛ bi<br />

⎞<br />

⎜1<br />

+ ⎟I ( 2-23)<br />

⎝ P ⎠<br />

Eq. ( 2-23) is also referred to as the Klink<strong>en</strong>berg correlation, where P is the mean<br />

pressure, k a is the appar<strong>en</strong>t gas permeability observed at the mean pressure, and k ∞ is<br />

called “true” permeability or Klink<strong>en</strong>berg permeability. For a large average pressure, the<br />

correction factor in par<strong>en</strong>theses goes to zero, and the appar<strong>en</strong>t and true permeabilities t<strong>en</strong>d<br />

to become equal. As the average pressure <strong>de</strong>creases, the two permeabilities can <strong>de</strong>viate<br />

significantly from each other. This behavior is confirmed by data pres<strong>en</strong>ted by<br />

Klink<strong>en</strong>berg (1941) for glass filters and core samples and by Reda (1987) for tuff.<br />

The gas slip factor b i is a coeffici<strong>en</strong>t that <strong>de</strong>p<strong>en</strong>ds on the mean free path of a particular gas<br />

and the average pore radius of the porous medium b i can be calculated by<br />

b i<br />

4 fλP<br />

= ( 2-24)<br />

l<br />

β<br />

where, l β is the radius of a capillary or a pore, λ is the mean free path of the gas<br />

molecules, and f is proportionality factor. The Klink<strong>en</strong>berg coeffici<strong>en</strong>t for air can be<br />

estimated as<br />

33


air<br />

b air<br />

= 0.<br />

11k<br />

l<br />

−<br />

= 0.<br />

86kl<br />

−0.<br />

39<br />

0.<br />

33<br />

, with 10<br />

, whit 10<br />

−14<br />

−14<br />

> l<br />

−19<br />

k > 10 Heid et al. (1950)<br />

> l<br />

−17<br />

k > 10 Jones and Ow<strong>en</strong>s (1980)<br />

34<br />

( 2-25)<br />

( 2-26)<br />

the Klink<strong>en</strong>berg coeffici<strong>en</strong>t for a giv<strong>en</strong> porous medium is differ<strong>en</strong>t for each gas and is<br />

<strong>de</strong>p<strong>en</strong>d<strong>en</strong>t on the local temperature. The Klink<strong>en</strong>berg factor can be corrected for differ<strong>en</strong>t<br />

conditions as follows [41]<br />

k<br />

g<br />

1 2<br />

⎛ b ⎛ ⎞⎛<br />

⎞ ⎛ ⎞<br />

i ⎞<br />

1<br />

⎜<br />

μ M<br />

i<br />

ref ⎟ ⎜<br />

Ti<br />

= k ⎜ ⎟ =<br />

⎟<br />

⎜<br />

⎟<br />

l + bi<br />

bref<br />

( 2-27)<br />

⎝ P ⎠ ⎜ ⎟ ⎜ ⎟<br />

⎝ μairTref<br />

⎠⎝<br />

M i ⎠ ⎝ Tref<br />

⎠<br />

where subscript ref refers to the refer<strong>en</strong>ce gas, which is usually air, and M is the<br />

molecular weight. The temperature T is in absolute units.<br />

Another transport mechanism, configurational diffusion, occurred in very low permeability<br />

of approximately<br />

−21<br />

2<br />

10 m<br />

1 2<br />

, where the gas molecule size is comparable to the pore<br />

diameter. The gas-phase permeability may be differ<strong>en</strong>t than the liquid-phase permeability<br />

due to this effect.<br />

2.5 Transi<strong>en</strong>t conduction and convection heat transport<br />

The process of volume averaging begins by forming the superficial average of Eqs. ( 2-1) to<br />

( 2-4), in the case of a homog<strong>en</strong>eous medium for the β-phase<br />

∂T<br />

∂t<br />

∂<br />

T<br />

∂t<br />

β ( ρ ) + ( ρc<br />

) ∇ T ) = ∇.(<br />

k ∇T<br />

)<br />

c p β<br />

p<br />

( ρc<br />

) + ( ρc<br />

)<br />

p<br />

1<br />

+<br />

V<br />

β<br />

∫<br />

n<br />

βσ<br />

Aβσ<br />

β<br />

. k ∇T<br />

dA<br />

β<br />

β<br />

p<br />

β<br />

β<br />

.( β v β<br />

β β , in the β-phase ( 2-28)<br />

∇.<br />

v<br />

β<br />

T<br />

β<br />

1<br />

+<br />

V<br />

∫<br />

n<br />

βσ<br />

Aβσ<br />

v T dA = ∇.<br />

k ∇T<br />

β<br />

β<br />

β<br />

β<br />

( 2-29)<br />

And, with the spatial <strong>de</strong>composition of the temperature and velocity for the β-phase<br />

~<br />

T = T − T<br />

β<br />

, ~ v = v − v<br />

β<br />

, ~ v<br />

~<br />

= 0 , T = 0<br />

( 2-30)<br />

β<br />

β<br />

we have<br />

β<br />

β<br />

β<br />

β<br />

β<br />

β


ε<br />

β<br />

∂ β<br />

( ρc<br />

) T + ( ρc<br />

)<br />

p<br />

β<br />

∂t<br />

⎛ ⎛<br />

= ∇.<br />

⎜<br />

k ⎜ T<br />

⎜ β ∇ β ⎜<br />

⎝ ⎝<br />

1<br />

+<br />

V ∫ n<br />

ε β<br />

∂<br />

p β ∂t<br />

Tβ<br />

p<br />

⎛ ⎛<br />

β<br />

= ∇<br />

⎜<br />

k ⎜<br />

1<br />

. β ε β∇<br />

Tβ<br />

+<br />

⎜ ⎜<br />

V<br />

⎝ ⎝<br />

β<br />

p<br />

βσ<br />

Aβσ<br />

β<br />

( ρc<br />

) + ( ρc<br />

)<br />

β<br />

∫<br />

β<br />

∇.<br />

ε<br />

β<br />

T<br />

β<br />

⎞⎞<br />

T dA⎟⎟<br />

1<br />

β +<br />

⎟⎟<br />

V<br />

⎠⎠<br />

∇.<br />

ε<br />

n<br />

βσ<br />

Aβσ<br />

β<br />

T<br />

β<br />

β<br />

β<br />

∫<br />

v<br />

n<br />

β<br />

βσ<br />

Aβσ<br />

v<br />

β<br />

β<br />

β<br />

. k ∇T<br />

dA<br />

+<br />

β<br />

35<br />

+<br />

~<br />

T ~ v<br />

~<br />

⎞ β ⎞<br />

( T T ) dA⎟⎟<br />

1<br />

β + β + nβσ.<br />

kβ∇Tβ<br />

dA<br />

⎟⎟<br />

⎠⎠<br />

β<br />

~<br />

T ~ v<br />

Imposing the l<strong>en</strong>gth-scale constraint to obtain the intrinsic form<br />

β<br />

V<br />

β<br />

β β<br />

∫<br />

Aβσ<br />

∂ β<br />

β<br />

β<br />

( ρc<br />

) T + ε ( ρc<br />

) v . ∇ T + ( ρc<br />

)<br />

β<br />

( 2-31)<br />

( 2-32)<br />

~<br />

ε β p<br />

β β p β<br />

β<br />

p ∇.<br />

T ~<br />

β v<br />

β<br />

β<br />

β<br />

β<br />

∂t<br />

( 2-33)<br />

⎛ ⎛<br />

⎞<br />

⎜<br />

β<br />

⎞<br />

k ⎜<br />

1 ~<br />

= ∇ ∇ T + ∫ T dA⎟⎟<br />

1<br />

. β ε β β n βσ β + ∫ n βσ . k β ∇Tβ<br />

dA<br />

⎜ ⎜<br />

V ⎟⎟<br />

V<br />

A<br />

A<br />

⎝ ⎝<br />

βσ ⎠⎠<br />

βσ<br />

In differ<strong>en</strong>tial equations like equation ( 2-33) we can clearly id<strong>en</strong>tify differ<strong>en</strong>t terms as<br />

• The terms involving area integrals of the unit normal multiplied by the spatial<br />

<strong>de</strong>viations reflect the tortuosity of the porous medium, since they are highly<br />

<strong>de</strong>p<strong>en</strong>d<strong>en</strong>t on the geometry of the interfacial region.<br />

• The volume integrals of the velocity <strong>de</strong>viations multiplied by the temperature<br />

<strong>de</strong>viations are responsible for what is commonly known as hydrodynamic<br />

dispersion.<br />

• The area integrals of the unit normal multiplied by the diffusive fluxes are the<br />

contributions from interfacial mass transport.<br />

The σ-phase transport equation is analogous to Eq. ( 2-33) without the convection term and<br />

we list the result as<br />

ε<br />

σ<br />

1<br />

+<br />

V<br />

( ρc<br />

)<br />

∫<br />

p<br />

n<br />

σ<br />

σβ<br />

Aβσ<br />

∂<br />

∂t<br />

T<br />

. k ∇T<br />

dA<br />

σ<br />

σ<br />

σ<br />

σ<br />

⎛<br />

= ∇.<br />

⎜<br />

k<br />

⎜<br />

⎝<br />

σ<br />

⎛<br />

⎜ε<br />

σ ∇ T<br />

⎜<br />

⎝<br />

σ<br />

σ<br />

1<br />

+<br />

V<br />

∫<br />

n<br />

σβ<br />

Aσβ<br />

~ ⎞⎞<br />

T dA⎟⎟<br />

σ ⎟⎟<br />

⎠⎠<br />

( 2-34)


2.5.1 One equation local thermal equilibrium<br />

Since we have neglected Dufour effect, the heat transfer problem may be solved<br />

in<strong>de</strong>p<strong>en</strong>d<strong>en</strong>tly from Eqs. ( 2-5) and ( 2-6). This question has received a lot of att<strong>en</strong>tion in the<br />

literature. The one-equation equilibrium mo<strong>de</strong>l consists of a single transport equation for<br />

both the σ and β-regions. Wh<strong>en</strong> the two temperatures in the two regions are close <strong>en</strong>ough,<br />

the transport equations that repres<strong>en</strong>t the two-equation mo<strong>de</strong>l can be ad<strong>de</strong>d to produce this<br />

mo<strong>de</strong>l. We mean that the principle of local-scale heat equilibrium is valid. The conditions<br />

for the validity of a one-equation conduction mo<strong>de</strong>l have be<strong>en</strong> investigated by Quintard et<br />

al. (1993). They have examined the process of transi<strong>en</strong>t heat conduction for a two-phase<br />

system in terms of the method of volume averaging. Using two equation mo<strong>de</strong>ls, they have<br />

explored the principle of local thermal equilibrium as a function of various parameters, in<br />

particular the conductivity ratio, micro-scale and macro-scale dim<strong>en</strong>sionless times and<br />

topology [79]. The one-equation equilibrium mo<strong>de</strong>l is obtained directly from the twoequation<br />

mo<strong>de</strong>l by imposing the constraints associated with local mass equilibrium.<br />

The local equilibrium mo<strong>de</strong>l obtained wh<strong>en</strong> there is a fast exchange betwe<strong>en</strong> the differ<strong>en</strong>t<br />

regions, is characterized by<br />

β<br />

β<br />

σ<br />

σ<br />

T = T = T<br />

( 2-35)<br />

If we accept this i<strong>de</strong>a, Eq. ( 2-33) and Eq. ( 2-34) can be ad<strong>de</strong>d to obtain:<br />

∂<br />

( ε β ( ρc<br />

p ) + εσ<br />

( ρc<br />

p ) ) T + ε β ( ρc<br />

p )<br />

= ∇.<br />

−<br />

( ε k + ε k ) ∇ T )<br />

β<br />

~ ( ρc<br />

) ~<br />

p ∇.<br />

Tβ<br />

v β<br />

β<br />

β<br />

β<br />

σ<br />

σ<br />

σ<br />

∂t<br />

⎛ k<br />

+ ∇.<br />

⎜<br />

⎜ V<br />

⎝<br />

β<br />

β<br />

β .<br />

β<br />

σ<br />

∫ nβσTβ<br />

dA + ∇.<br />

⎟ ⎜ V<br />

A<br />

A<br />

βσ<br />

~<br />

v<br />

⎞<br />

⎟<br />

⎠<br />

∇ T<br />

36<br />

⎛<br />

⎜ k<br />

⎝<br />

∫<br />

βσ<br />

n<br />

σβ<br />

~ ⎞<br />

Tσ<br />

dA⎟<br />

⎟<br />

⎠<br />

( 2-36)<br />

Equation ( 2-33) is not too useful in its curr<strong>en</strong>t form because of the terms containing the<br />

spatial <strong>de</strong>viationsT β<br />

~ . Therefore, one seeks to relate this spatial <strong>de</strong>viation to the averaged<br />

temperature<br />

β<br />

β<br />

T and their gradi<strong>en</strong>t. This will help us to obtain a closure of the problem,<br />

i.e., to have <strong>en</strong>ough equations to allow a solution for the averaged temperature.<br />

In or<strong>de</strong>r to <strong>de</strong>velop this closure scheme, we <strong>de</strong>rive governing differ<strong>en</strong>tial equations for the<br />

spatial <strong>de</strong>viation by subtracting the average equation ( 2-33) from the point equation ( 2-1).<br />

We th<strong>en</strong> make a <strong>de</strong>termination of the most important terms in the governing equations for<br />

T β<br />

~ by using estimates of the or<strong>de</strong>r of magnitu<strong>de</strong> of all the terms.


Finally, we postulate the functional <strong>de</strong>p<strong>en</strong>d<strong>en</strong>ce of T β<br />

~ on<br />

37<br />

β<br />

β<br />

T by analyzing the form of<br />

the differ<strong>en</strong>tial equation and boundary conditions. The resulting constitutive equations will<br />

have functions that need to be evaluated in or<strong>de</strong>r to allow calculations of the important<br />

transport coeffici<strong>en</strong>ts.<br />

In or<strong>de</strong>r to <strong>de</strong>velop the governing differ<strong>en</strong>tial equation for T β<br />

~ we divi<strong>de</strong> Eq. ( 2-33) by ε β<br />

and the result can be expressed as<br />

∂ β<br />

( ρc<br />

) T + ( ρc<br />

)<br />

p<br />

β ∂t<br />

⎛ ⎛<br />

= ∇.<br />

⎜<br />

⎜<br />

k<br />

⎜<br />

β ⎜<br />

⎝ ⎝<br />

−1<br />

ε β<br />

+ ∫ n<br />

V<br />

β<br />

β ( ∇.<br />

Tβ<br />

)<br />

βσ<br />

Aβσ<br />

β<br />

+ ε<br />

β<br />

p<br />

β<br />

−1<br />

β<br />

. k ∇T<br />

dA −<br />

v<br />

β<br />

β<br />

β<br />

∇ε<br />

∇ T<br />

−1<br />

~<br />

( ρc<br />

) ε ~<br />

p β ∇.<br />

Tβ<br />

v β<br />

β<br />

. ∇ T<br />

β<br />

β<br />

β<br />

β<br />

⎛<br />

⎜<br />

ε<br />

+<br />

⎜ V<br />

⎝<br />

−1<br />

β<br />

∫ n βσ<br />

Aβσ<br />

~ ⎞⎞<br />

⎞<br />

Tβ<br />

dA⎟⎟<br />

⎟<br />

⎟⎟<br />

⎟<br />

⎠⎠<br />

⎠<br />

( 2-37)<br />

If we subtract Eq. ( 2-37) from Eq. ( 2-1), we obtain the following governing differ<strong>en</strong>tial<br />

equation for the spatial <strong>de</strong>viation temperature, T β<br />

~ in β-phase<br />

ρc<br />

p<br />

~<br />

∂Tβ<br />

+<br />

β ∂t<br />

ρc<br />

p v β . ∇Tβ<br />

− ρc<br />

β<br />

p v<br />

β β<br />

β<br />

. ∇ Tβ<br />

β<br />

= ∇.<br />

k β ∇<br />

β ⎛<br />

1<br />

1 k β ~ ⎞<br />

−<br />

−<br />

− ε β ∇ε<br />

β . k β ∇ Tβ<br />

− ε β ∇.<br />

⎜ n βσTβ<br />

dA⎟<br />

⎜ V ∫ ⎟<br />

⎝ Aβσ<br />

⎠<br />

−1<br />

ε β<br />

−<br />

V<br />

−1<br />

~<br />

n βσ . k β ∇Tβ<br />

dA − ( ρc<br />

) ε ~<br />

∫<br />

p β β ∇.<br />

Tβ<br />

v β<br />

~<br />

( ) ( ) ( ) ( T )<br />

Aβσ<br />

One can express the interfacial flux as<br />

1<br />

V<br />

∫<br />

n<br />

βσ<br />

Aβσ<br />

. k ∇T<br />

dA = −<br />

β<br />

β<br />

β 1 ~<br />

( ∇ε<br />

β ) . kβ<br />

∇ Tβ<br />

+ ∫ nβσ.<br />

kβ<br />

∇Tβ<br />

dA<br />

V<br />

Aβσ<br />

β<br />

( 2-38)<br />

( 2-39)<br />

In this equation we have ma<strong>de</strong> use of a theorem <strong>de</strong>veloped by Gray (1975), relating area<br />

integrals of the unit normal to gradi<strong>en</strong>ts in volume fraction, for this case<br />

1<br />

V<br />

∫ n βσdA<br />

Aβσ<br />

= −∇ε<br />

β<br />

( 2-40)<br />

Since the volume fraction of the β–phase have be<strong>en</strong> tak<strong>en</strong> as constant, this integral sum to<br />

zero.


The point temperature field T β will vary microscopically within each phase over distances<br />

on the or<strong>de</strong>r of the characteristic l<strong>en</strong>gth l β indicated in Fig. 2-1. This is also the<br />

characteristic l<strong>en</strong>gth associated with large variations in the spatial <strong>de</strong>viation field T β<br />

~ .<br />

However, the average field<br />

β<br />

β<br />

T is treated as being constant within the averaging<br />

volume,V . It un<strong>de</strong>rgoes significant variations only over distances L which is much<br />

greater than the characteristic l<strong>en</strong>gth l β . These two wi<strong>de</strong>ly differ<strong>en</strong>t l<strong>en</strong>gth scales in the<br />

problem helps us to simplify the transport equations for the spatial <strong>de</strong>viations by making<br />

or<strong>de</strong>r of magnitu<strong>de</strong> estimates of the terms in equation ( 2-38) and the equation for T β<br />

~ .<br />

The or<strong>de</strong>r of magnitu<strong>de</strong> of the non-local convective transport term can be expressed as<br />

~<br />

∇.<br />

T ~<br />

β v β<br />

⎛<br />

⎜<br />

= O<br />

⎜<br />

⎝<br />

β ~<br />

v T ⎞<br />

β β ⎟<br />

L ⎟<br />

⎠<br />

( 2-41)<br />

and the or<strong>de</strong>r of magnitu<strong>de</strong> of the local convective transport is<br />

⎛<br />

~ ⎜<br />

v β . ∇Tβ<br />

= O⎜<br />

⎜<br />

⎝<br />

β ~<br />

v ⎞<br />

β Tβ<br />

⎟<br />

l<br />

⎟<br />

β ⎟<br />

⎠<br />

( 2-42)<br />

This indicates that the non-local convective transport can be neglected wh<strong>en</strong>-ever<br />

l β


⎛<br />

⎜ k ~<br />

. βσ<br />

⎜<br />

p ⎝ β Aβσ<br />

⎞<br />

( ) ( ) ⎟ ⎟<br />

β<br />

⎟ > 1<br />

( 2-50)<br />

Un<strong>de</strong>r this conditions the quasi-steady approximation can be ma<strong>de</strong> in equation ( 2-47)<br />

~<br />

∂Tβ<br />


~<br />

~<br />

BC2: − ∇T<br />

= −n<br />

. k ∇T<br />

+ n . ( k − k )<br />

β<br />

n βσ.<br />

k β β βσ σ σ βσ β σ ∇ Tβ<br />

, at A ( 2-54)<br />

βσ<br />

~<br />

~<br />

T β = f r,<br />

t , at Aβe & T σ = g(<br />

r,<br />

t)<br />

( 2-55)<br />

BC3: ( )<br />

0 = ∇.<br />

−1<br />

~ ε β ~<br />

( k ∇T<br />

) − n . k T dA<br />

β<br />

β<br />

V<br />

I. Closure variable<br />

∫<br />

βσ<br />

Aβσ<br />

β<br />

β<br />

40<br />

( 2-56)<br />

In or<strong>de</strong>r to obtain a closure, we need to relate the spatial <strong>de</strong>viations T β<br />

~ , T σ<br />

~ to the average<br />

temperature T . The transport equations for spatial <strong>de</strong>viation fields are linear in the<br />

averaged terms. We are thus <strong>en</strong>couraged to look for linear relations betwe<strong>en</strong> the spatial<br />

<strong>de</strong>viations and average conc<strong>en</strong>trations of the type [14]<br />

~<br />

Tβ = b Tβ . ∇ T<br />

( 2-57)<br />

~<br />

Tσ = b Tσ . ∇ T<br />

( 2-58)<br />

in which b Tβ<br />

and b Tσ<br />

are referred to as the closure variables for solid and liquid<br />

respectively. These vectors are functions of position only, since the time <strong>de</strong>p<strong>en</strong>d<strong>en</strong>ce of the<br />

T β<br />

~ and T σ<br />

~ comes only from the time <strong>de</strong>p<strong>en</strong>d<strong>en</strong>ce of the average temperature appearing in<br />

the equations and boundary conditions for the spatial <strong>de</strong>viation. If we substitute equations<br />

( 2-57) and ( 2-58) into equation ( 2-52) to ( 2-56), we can <strong>de</strong>rive transport equations for the<br />

closure functions for each phases. In doing this we can neglect higher <strong>de</strong>rivatives of the<br />

average fields in the expressions for the gradi<strong>en</strong>ts. This approximation is consist<strong>en</strong>t with<br />

the constraint l


− ε<br />

V<br />

−1<br />

σ<br />

∫<br />

Aβσ<br />

n<br />

βσ<br />

2<br />

k ∇b<br />

dA = k ∇ b<br />

β<br />

Tβ<br />

Periodicity: ( r ) = b ( r)<br />

σ<br />

Tσ<br />

41<br />

( 2-62)<br />

bT β + l i Tβ<br />

& b ( r + l ) = b σ ( r)<br />

, i=1,2,3 ( 2-63)<br />

β<br />

T σ i T<br />

Averages: b = 0 , b = 0<br />

( 2-64)<br />

Tβ<br />

σ<br />

Tσ<br />

This way of writing the problem, i.e., un<strong>de</strong>r an integro-differ<strong>en</strong>tial form, is reminisc<strong>en</strong>t of<br />

the fact that this must be compatible with the full two-equation mo<strong>de</strong>l as <strong>de</strong>scribed, for<br />

instance, in [82]. However, following the mathematical treatm<strong>en</strong>t <strong>de</strong>scribed also in this<br />

paper (using the <strong>de</strong>composition <strong>de</strong>scribed by Eqs. 20 in ref [82]), this problem reduces to<br />

(the proof involves the use of periodicity conditions) the following problem.<br />

Problem I:<br />

2<br />

( c ) vβ<br />

. b β + ( ρc<br />

) ~ vβ<br />

= kβ∇<br />

b β<br />

ρ ∇ T<br />

T<br />

( 2-65)<br />

p β<br />

p β<br />

BC1: b Tβ<br />

= bTσ<br />

, at A βσ<br />

( 2-66)<br />

BC2: − . ∇b<br />

= −n<br />

. k ∇b<br />

+ n . ( k − k )<br />

n βσ kβ Tβ<br />

βσ σ Tσ<br />

βσ β σ , at βσ<br />

A ( 2-67)<br />

kσ bTσ<br />

2<br />

0 = ∇<br />

( 2-68)<br />

Periodicity: bT β ( r + l i ) = bTβ<br />

( r)<br />

& b ( r + l ) = b σ ( r)<br />

, i=1,2,3 ( 2-69)<br />

β σ<br />

T β Tβ σ Tσ<br />

T σ i T<br />

Averages: b = ε b + ε b = 0<br />

( 2-70)<br />

In fact, the resulting field is also compatible with Eqs. ( 2-64), which is consist<strong>en</strong>t with the<br />

local-equilibrium closure being compatible with the one from the two-equation mo<strong>de</strong>l, as a<br />

limit case.<br />

II. Transport equation for averaged temperature<br />

In or<strong>de</strong>r to obtain the closed form of the macroscopic equation, we recall Eq. ( 2-36)<br />

∂<br />

( ε β ( ρc<br />

p ) + ε σ ( ρc<br />

p ) ) T + ε β ( ρc<br />

p )<br />

= ∇.<br />

−<br />

( ε k + ε k ) ∇ T )<br />

β<br />

~ ( ρc<br />

) ~<br />

p ∇.<br />

Tβ<br />

v β<br />

β<br />

β<br />

β<br />

σ<br />

σ<br />

σ<br />

∂t<br />

⎛ k<br />

+ ∇.<br />

⎜<br />

⎜ V<br />

⎝<br />

β<br />

~<br />

v<br />

β<br />

⎞<br />

⎟<br />

⎠<br />

β<br />

σ<br />

∫ n βσTβ<br />

dA + ∇.<br />

n σβ<br />

⎟ ⎜ V ∫<br />

Aβσ<br />

Aβσ<br />

β<br />

. ∇ T<br />

⎛<br />

⎜ k<br />

⎝<br />

~ ⎞<br />

Tσ<br />

dA⎟<br />

⎟<br />

⎠<br />

( 2-71)


In or<strong>de</strong>r to obtain a transport equation for the averaged temperature T , we substitute the<br />

repres<strong>en</strong>tations for T β<br />

~ and T σ<br />

~ (equations ( 2-57) and ( 2-58)) into the spatially averaged<br />

convective diffusion equation for two phases, equation ( 2-36) or ( 2-71). Note that, in this<br />

case, one cannot neglect terms involving second <strong>de</strong>rivatives of the average conc<strong>en</strong>tration.<br />

It is a mistake to neglect these terms in the transport equations since they are of the same<br />

or<strong>de</strong>r as the tortuosity or dispersion t<strong>en</strong>sors. As was done in the <strong>de</strong>velopm<strong>en</strong>t of the<br />

equations for the average temperature, we treat all averaged quantities as constants within<br />

the averaging volume. Therefore, the transport equation obtained has the form<br />

∂<br />

β<br />

( ε β ( ρc<br />

p ) + ε σ ( ρc<br />

p ) ) T + ( ρc<br />

p ) ∇.<br />

( ε β v β T )<br />

β<br />

σ ∂t<br />

⎛⎛<br />

⎜⎜<br />

k β − k<br />

= ∇.<br />

( ε + ) ) +<br />

⎜ β λβ<br />

ε σ λσ<br />

I<br />

⎜<br />

V<br />

⎝⎝<br />

−<br />

( ρc<br />

) ∇.<br />

( ~ v b . ∇ T )<br />

p<br />

β<br />

β Tβ<br />

∫<br />

β<br />

σ<br />

n βσ<br />

Aβσ<br />

b<br />

Tβ<br />

⎞<br />

dA⎟.<br />

∇ T<br />

⎟<br />

⎠<br />

42<br />

⎞<br />

⎟<br />

⎟<br />

⎠<br />

( 2-72)<br />

A product (of two vectors) such as n b β is called a dyad product and is a special form<br />

βσ T<br />

of the second-or<strong>de</strong>r t<strong>en</strong>sors. Each t<strong>en</strong>sor is <strong>de</strong>composed as<br />

n βσ = n s + n s + nβσ<br />

s<br />

( 2-73)<br />

βσ 1<br />

1<br />

βσ 2<br />

2<br />

3<br />

3<br />

where s i is the unit vector in the i-direction. Th<strong>en</strong> the dyad product is giv<strong>en</strong> by<br />

⎡nβσ<br />

1bTβ<br />

1 nβσ<br />

1bTβ<br />

2 nβσ<br />

1bTβ<br />

3 ⎤<br />

⎢<br />

⎥<br />

n βσbTβ<br />

= ⎢nβσ<br />

2bTβ<br />

1 nβσ<br />

2bTβ<br />

2 nβσ<br />

2bTβ<br />

3⎥<br />

( 2-74)<br />

⎢<br />

⎥<br />

⎣nβσ<br />

3bTβ<br />

1 nβσ<br />

3bTβ<br />

2 nβσ<br />

3bTβ<br />

3⎦<br />

Also, the unit t<strong>en</strong>sor used in Eq. ( 2-77) is<br />

⎡1<br />

0 0⎤<br />

I =<br />

⎢ ⎥<br />

⎢<br />

0 1 0<br />

⎥<br />

( 2-75)<br />

⎢⎣<br />

0 0 1⎥⎦<br />

in Cartesian coordinates.<br />

Th<strong>en</strong> the closed form of the convective-dispersion governing equation for T can be<br />

writt<strong>en</strong><br />

∂<br />

*<br />

( ( cp ) + ( cp<br />

) ) T + ( c p ) ∇.<br />

( v T ) = ∇.<br />

( k . ∇ T )<br />

β<br />

β ρ εσ<br />

ρ<br />

ρ ε β β<br />

ε ( 2-76)<br />

β<br />

σ<br />

β<br />

∂t<br />

where<br />

*<br />

k is the thermal dispersion t<strong>en</strong>sor giv<strong>en</strong> by


( ε + ε k )<br />

( k − k )<br />

( ρcp<br />

) vβ<br />

Tβ<br />

*<br />

k = +<br />

− ~<br />

∫<br />

β σ<br />

βk<br />

β σ σ I<br />

nβσbTβ<br />

dA b<br />

β<br />

V A<br />

βσ<br />

43<br />

( 2-77)<br />

As an illustration of such a local-equilibrium situation, we will compare a direct simulation<br />

of the pore-scale equations with a macro-scale prediction. The geometry is an array of NUC<br />

of the periodic Unit Cell (UC) shown in Fig. 2-6 . The initial temperature in the domain is<br />

a constant, TC . The fluid is injected at x=0 at temperature TH. The temperature is imposed<br />

at the exit boundary and is equal to TC. This latter boundary condition has be<strong>en</strong> tak<strong>en</strong> for a<br />

practical reason: we have ongoing experim<strong>en</strong>ts using the two-bulb method, which is<br />

closely <strong>de</strong>scribed by this kind of boundary-value problem. In addition, this particular<br />

problem will help us to illustrate some theoretical consi<strong>de</strong>rations giv<strong>en</strong> below. The<br />

parameters <strong>de</strong>scribing the case were:<br />

N k k<br />

UC = 120 ; lUC<br />

=<br />

−3<br />

2× 10 m; β = 1W/m.K ; σ = 4 W/m.K ; ε β = 0.615<br />

p β = ×<br />

6 3<br />

p σ = ×<br />

6 3<br />

β<br />

β<br />

= ×<br />

−4<br />

*<br />

( ρc ) 4.18 10 J/m K ; ( ρc<br />

) 4 10 J/m K ; v 1.4 10 m / s<br />

k<br />

= 1.607 W/m.K<br />

( 2-78)<br />

An example of comparison betwe<strong>en</strong> the averaged temperatures obtained from the direct<br />

simulation and theoretical predictions is giv<strong>en</strong> Fig. 2-2. We consi<strong>de</strong>red three stages:<br />

• a short time after injection, which is oft<strong>en</strong> the source of a discrepancy betwe<strong>en</strong><br />

actual fields and macro-scale predicted ones, because of the vicinity of the<br />

boundary,<br />

• an intermediate time, i.e., a field less impacted by boundary conditions,<br />

• a long time typical of the steady-state condition associated to the initial boundaryvalue<br />

problem un<strong>de</strong>r consi<strong>de</strong>ration.


Fig. 2-2. Normalized temperature versus position, for three differ<strong>en</strong>t times (triangle, Direct Numerical<br />

β<br />

σ<br />

Simulation= ( T −T<br />

) ( T −T<br />

) ; circles, Direct Numerical Simulation = ( T −T<br />

) ( T −T<br />

)<br />

β C H C<br />

σ C H C<br />

Local-equilibrium mo<strong>de</strong>l= ( T −<br />

T ) ( T −T<br />

)<br />

C<br />

H<br />

C<br />

44<br />

; solid line,


We see on these figures a very good agreem<strong>en</strong>t betwe<strong>en</strong> the direct simulations and the<br />

predictions with the local equilibrium mo<strong>de</strong>l. This illustrates the fact that the local-<br />

equilibrium mo<strong>de</strong>l does allow to repres<strong>en</strong>t correctly the system behaviour for mo<strong>de</strong>rate<br />

contrasts of the pore-scale physical properties. What happ<strong>en</strong>s wh<strong>en</strong> this contrast becomes<br />

dramatic, i.e., wh<strong>en</strong> the pore-scale characteristic times are very differ<strong>en</strong>t? To illustrate the<br />

problem, we <strong>de</strong>signed such a case by taking the following parameters:<br />

N k k<br />

UC = 480 ; lUC<br />

=<br />

−3<br />

2× 10 m; β = 1W/mK ; σ = 0.01W/mK ; ε β = 0.615<br />

cp β = ×<br />

6 3<br />

cp σ = ×<br />

6 3<br />

vβ β<br />

= ×<br />

−5<br />

s<br />

*<br />

( ρ ) 4 10 J/m K ; ( ρ ) 4 10 J/m K ; 6.95 10 m/<br />

k<br />

= 0.455W/m.K<br />

45<br />

( 2-79)<br />

The comparison betwe<strong>en</strong> the averaged temperatures obtained from direct numerical<br />

simulations and the theoretical predictions of the local-equilibrium mo<strong>de</strong>l are pres<strong>en</strong>ted in<br />

Fig. 2-3 for three differ<strong>en</strong>t times. At early stages, we see a clear differ<strong>en</strong>ce betwe<strong>en</strong> the<br />

averaged temperatures of the two phases, and also a clear differ<strong>en</strong>ce with the localequilibrium<br />

predictions. This differ<strong>en</strong>ce is also visible for intermediate times, and one sees<br />

that the local-equilibrium mo<strong>de</strong>l has an effective conductivity which is too small.<br />

However, at steady-state, it is remarkable to see that the temperature fields revert to the<br />

local-equilibrium conditions and that, <strong>de</strong>spite the steep gradi<strong>en</strong>t near the boundary, the<br />

local-equilibrium mo<strong>de</strong>ls offers a very good prediction. It must be pointed out that this<br />

possibility has not be<strong>en</strong> docum<strong>en</strong>ted in the literature, and this may explain certain<br />

confusion in the discussion about the various macro-scale mo<strong>de</strong>ls. Without going into<br />

many <strong>de</strong>tails, we may summarize the discussion as follows:<br />

• for mo<strong>de</strong>rate thermal properties contrasts, the local-equilibrium predictions are<br />

very good, and not very s<strong>en</strong>sitive to boundary conditions or initial conditions,<br />

• the situation is much more complex for higher contrasts, which lead to nonequilibrium<br />

conditions.


Fig. 2-3. Normalized temperature versus position, for three differ<strong>en</strong>t times (triangle, Direct Numerical<br />

β<br />

σ<br />

Simulation= ( T −T<br />

) ( T −T<br />

) ; circles, Direct Numerical Simulation = ( T −T<br />

) ( T −T<br />

)<br />

β C H C<br />

σ C H C<br />

Local-equilibrium mo<strong>de</strong>l= ( T −<br />

T ) ( T −T<br />

)<br />

C<br />

H<br />

C<br />

46<br />

; solid line,


If the local equilibrium assumption does not hold, the differ<strong>en</strong>t stages for the typical<br />

problem consi<strong>de</strong>red here are as following:<br />

• early stages: initial conditions with sharp gradi<strong>en</strong>ts and the vicinity of boundaries<br />

create non-equilibrium situations that are difficult to homog<strong>en</strong>ize. They may be<br />

mo<strong>de</strong>lled through modified boundary conditions ([71], [16]), mixed mo<strong>de</strong>ls (i.e., a<br />

small domain keeping pore-scale <strong>de</strong>scription such as in [6]).<br />

• two-equation behaviour: in g<strong>en</strong>eral, the initial sharp gradi<strong>en</strong>ts are smoothed after<br />

some time and more homog<strong>en</strong>izable conditions are found. Differ<strong>en</strong>t mo<strong>de</strong>ls may<br />

be used: mixed mo<strong>de</strong>ls, differ<strong>en</strong>t types of two-equation mo<strong>de</strong>ls (see a review and<br />

discussion in [83]), or more sophisticated equations in [105]. Two-equation mo<strong>de</strong>ls<br />

may be more or less sophisticated, for instance, two-equation mo<strong>de</strong>ls with first<br />

or<strong>de</strong>r exchange terms [11, 82, 79, 115, 116] or two-equation mo<strong>de</strong>ls with more<br />

elaborate exchange terms like convolution terms that would mo<strong>de</strong>l non local and<br />

memory effects [64]. This is beyond the scope of this paper to <strong>de</strong>velop such a<br />

theory for our double-diffusion problem.<br />

• asymptotic behaviour: if the medium has an infinite ext<strong>en</strong>t (this can also be<br />

mimicked by convective conditions at the exit for a suffici<strong>en</strong>tly large domain),<br />

cross diffusion may lead to a so-called asymptotic behaviour which may be<br />

<strong>de</strong>scribed by a one-equation mo<strong>de</strong>l with a differ<strong>en</strong>t effective thermal conductivity,<br />

larger than the local-equilibrium value. This asymptotic behaviour for dispersion<br />

problems has be<strong>en</strong> investigated by several authors and the link betwe<strong>en</strong> the oneequation<br />

mo<strong>de</strong>l obtained and the properties of the two-equation mo<strong>de</strong>l well<br />

docum<strong>en</strong>ted ([116], [2], [81]). The one-equation non-equilibrium mo<strong>de</strong>l may be<br />

<strong>de</strong>rived directly by a proper choice of the averaged conc<strong>en</strong>tration/temperature and<br />

<strong>de</strong>viations as in [81] and [65].<br />

• Complex history: It must be emphasized that non-equilibrium mo<strong>de</strong>ls<br />

corresponding to the asymptotic behaviour require special situations to be valid. If<br />

ev<strong>en</strong>ts along the flow path change due to forcing terms like source terms,<br />

heterog<strong>en</strong>eities, boundaries, the conditions leading to the asymptotic behaviour are<br />

disturbed and a differ<strong>en</strong>t history <strong>de</strong>velops. This is what happ<strong>en</strong>ed in our test case.<br />

The boundary effects damp<strong>en</strong>ed the asymptotic behaviour that has probably tak<strong>en</strong><br />

place in our system (in the abs<strong>en</strong>ce of an interpretation with two-equation mo<strong>de</strong>ls<br />

or one-equation asymptotic mo<strong>de</strong>ls, we cannot distinguish betwe<strong>en</strong> the two<br />

47


possibilities, while the large ext<strong>en</strong>t of the domain has probably favoured an<br />

asymptotic behaviour) and this led to a steady-state situation well <strong>de</strong>scribed by the<br />

local-equilibrium mo<strong>de</strong>l. This possibility has not be<strong>en</strong> se<strong>en</strong> by many investigators.<br />

However, it must be tak<strong>en</strong> into account for practical applications. H<strong>en</strong>ce, for our<br />

test case, it would be better to use a two-equation mo<strong>de</strong>l, which truly embeds the<br />

one-equation local-equilibrium mo<strong>de</strong>l, than the asymptotic mo<strong>de</strong>l that would fail to<br />

catch the whole history.<br />

Now we have at our disposal a mapping vector that gives the local temperature field in<br />

terms of the averaged value. It is important to remark that the upscaling of the heat<br />

equation problem has be<strong>en</strong> solved in<strong>de</strong>p<strong>en</strong>d<strong>en</strong>tly from the solute transport problem. This<br />

feature is a key approximation that will simplify the treatm<strong>en</strong>t of the solute transport<br />

equation as explained in section 2.6.<br />

2.5.2 Two equation mo<strong>de</strong>l<br />

In this case the time and l<strong>en</strong>gth scales are such that a unique macroscopic or effective<br />

medium cannot repres<strong>en</strong>t the macroscopic behavior of the two phases. The two-equation<br />

mo<strong>de</strong>l consists of separate heat transport equations for both the σ and β-phases. The<br />

dominant coupling betwe<strong>en</strong> the two equations is repres<strong>en</strong>ted by an inter-phase flux that<br />

<strong>de</strong>p<strong>en</strong>ds, in the simpler version, on an exchange coeffici<strong>en</strong>t and the differ<strong>en</strong>ce betwe<strong>en</strong> the<br />

temperatures of the two phases. The spatial <strong>de</strong>viation temperatures in terms of the<br />

macroscopic source terms is oft<strong>en</strong> repres<strong>en</strong>ted as<br />

~<br />

β<br />

σ<br />

β σ<br />

= b . ∇ T + b . ∇ T − s T − T<br />

( 2-80)<br />

Tβ Tββ<br />

β Tβσ<br />

~<br />

β<br />

Tσ Tσβ<br />

. ∇ Tβ<br />

+ bTσσ<br />

σ<br />

σ<br />

σ<br />

β<br />

σ<br />

( β<br />

σ )<br />

( Tβ<br />

β<br />

− Tσ<br />

σ )<br />

= b . ∇ T − s<br />

( 2-81)<br />

Here the mapping vectors and the scalars are obtained from the solution of steady, pore<br />

scale closure problems (see Quintard et al. 1997). Th<strong>en</strong>, the macroscopic equation for the<br />

β-phase, is expressed as [82]<br />

ε<br />

β<br />

= ∇<br />

∂ β<br />

β β<br />

( ρc<br />

p ) Tβ<br />

+ ( ρc<br />

) ( )<br />

β<br />

p ∇.<br />

ε<br />

β β β Tβ<br />

− u ββ.<br />

∂t<br />

β<br />

σ<br />

β σ<br />

. ( k ββ . ∇ Tβ<br />

+ k βσ . ∇ Tσ<br />

) − avh(<br />

Tβ<br />

− Tσ<br />

)<br />

v ∇ T − u βσ . ∇ T<br />

48<br />

β<br />

β<br />

σ<br />

σ<br />

( 2-82)<br />

In this equation, the effective properties such as k ββ , k βσ , u ββ , u βσ and the volumetric<br />

heat exchange coeffici<strong>en</strong>t av h , are obtained explicitly from the mapping vectors, and


epres<strong>en</strong>tation of these coeffici<strong>en</strong>ts are giv<strong>en</strong> in [82]. An equation analogous to equation<br />

( 2-82) <strong>de</strong>scribes the intrinsic average temperature for the σ-phase, and this equation is<br />

giv<strong>en</strong> by<br />

ε<br />

( c )<br />

σ ρ p<br />

= ∇.<br />

σ<br />

∂<br />

∂t<br />

T<br />

σ<br />

σ<br />

− u<br />

σβ<br />

. ∇ T<br />

σ<br />

σ<br />

β<br />

σ<br />

β σ<br />

( k σβ . ∇ Tβ<br />

+ k σσ . ∇ Tσ<br />

) + avh(<br />

Tβ<br />

− Tσ<br />

)<br />

β<br />

β<br />

− u<br />

σσ<br />

. ∇ T<br />

49<br />

( 2-83)<br />

This two-equations mo<strong>de</strong>l is fully compatible with the one-equation mo<strong>de</strong>l with the<br />

effective thermal dispersion t<strong>en</strong>sor of equation ( 2-76) giv<strong>en</strong> by [82]<br />

k *<br />

= k + k + k + k<br />

( 2-84)<br />

ββ<br />

βσ<br />

σβ<br />

2.5.3 Non-equilibrium one-equation mo<strong>de</strong>l<br />

σσ<br />

If the local equilibrium assumption is not ma<strong>de</strong>, it is possible to obtain a one-equation nonequilibrium<br />

mo<strong>de</strong>l which consists of a single transport equation for both the σ and βregions.<br />

Similarly, it can be shown that, the two equation mo<strong>de</strong>l <strong>de</strong>scribed in the last<br />

section reduces to a single dispersion equation for suffici<strong>en</strong>tly long time. It can be obtained<br />

by <strong>de</strong>fining an <strong>en</strong>thalpy averaged temperature and working with the upscaling process by<br />

<strong>de</strong>fining <strong>de</strong>viations with respect to this temperature [65]. One can begin with the twoequation<br />

mo<strong>de</strong>l, <strong>de</strong>termine the sum of the spatial mom<strong>en</strong>ts of the two equations, and<br />

construct a one-equation mo<strong>de</strong>l that matches the sum of the first three spatial mom<strong>en</strong>ts in<br />

the long-time limit. The second analysis yields exactly the same equation as the first as<br />

explained in Quintard et al. (2001) for the case of dispersion in heterog<strong>en</strong>eous systems<br />

[81]. A complete three-dim<strong>en</strong>sional mom<strong>en</strong>t’s analysis associated with a two-equation<br />

mo<strong>de</strong>l has be<strong>en</strong> proposed in refer<strong>en</strong>ce [115]. It is shown that a mo<strong>de</strong>l with two equations<br />

converges asymptotically to a mo<strong>de</strong>l with one equation, and it is possible to obtain an<br />

expression for the asymptotic global dispersion coeffici<strong>en</strong>t. A similar analysis was<br />

pres<strong>en</strong>ted in the case of miscible transport in a stratified structure [2]; in this case some<br />

coeffici<strong>en</strong>ts are zero.<br />

The processes of a spatial mom<strong>en</strong>t analysis are listed in Table 2-1.<br />

Table 2-1. Objectives of each or<strong>de</strong>r of mom<strong>en</strong>tum analysis<br />

Or<strong>de</strong>r of mom<strong>en</strong>t Definition<br />

Zeroth The total amount of field pres<strong>en</strong>t in each phase


First The average position<br />

Second Measure of the spread of the pulse relative to its average position<br />

Consi<strong>de</strong>r a pulse introduced into spatially infinite system at time t=0. As the pulse<br />

transported, the temperature in each phase will change with position and time according to<br />

Eqs. ( 2-82) and ( 2-83).<br />

From zeroth spatial mom<strong>en</strong>t, we can see that a quasi-equilibrium condition is reached<br />

wh<strong>en</strong><br />

( ρc<br />

p ) εσ<br />

( ρc<br />

)<br />

β p σ<br />

( ρc<br />

) + ε ( ρc<br />

)<br />

ε<br />

t >><br />

a h<br />

( 2-85)<br />

v<br />

β<br />

( ε β p σ p )<br />

β<br />

σ<br />

The first or<strong>de</strong>r mom<strong>en</strong>t provi<strong>de</strong>s that the differ<strong>en</strong>ce betwe<strong>en</strong> the two mean pulse positions<br />

is a constant and, as a result, both pulses will move at the same velocity. Giv<strong>en</strong> this result,<br />

one tries to obtain the rate of spread of the pulses in each phase relative to their mean<br />

position by second mom<strong>en</strong>t analysis which shows that the differ<strong>en</strong>ce in the pulse spreads is<br />

constant. Th<strong>en</strong>, consi<strong>de</strong>ring a flow parallel to the x-axis, the one-equation mo<strong>de</strong>l can be<br />

writt<strong>en</strong> as<br />

β<br />

2<br />

∂<br />

∂<br />

( ( ) ( ) ) ( ε β vβ<br />

T ) ∂<br />

( )<br />

( ε T )<br />

∞ * β ∞<br />

ε β ρc<br />

p + εσ<br />

ρc<br />

p T + ρcp<br />

= k∞<br />

β<br />

σ<br />

∂t<br />

The asymptotic thermal dispersion coeffici<strong>en</strong>t<br />

∞<br />

β<br />

∂x<br />

β<br />

2<br />

β ( ρc<br />

p ) v<br />

β β ⎟<br />

⎞<br />

⎠<br />

( ρc<br />

) + ε ( ρc<br />

)<br />

( ) 2<br />

ε β p σ p<br />

β<br />

50<br />

∂x<br />

*<br />

k ∞ is giv<strong>en</strong> by [115, 2]<br />

σ<br />

( 2-86)<br />

⎜<br />

⎛ε<br />

*<br />

k = k ββ + k βσ + kσβ<br />

+ k<br />

⎝<br />

σσ +<br />

( 2-87)<br />

∞<br />

a h<br />

*<br />

k ∞ can be much greater than<br />

v<br />

*<br />

k as it is illustrated by numerical examples obtained for the<br />

case of a stratified system in Ahmadi et al. (1998). We note that, in this case there is no<br />

reason for equality betwe<strong>en</strong> the regional averages; however, the differ<strong>en</strong>ce betwe<strong>en</strong> the<br />

two regional temperatures will g<strong>en</strong>erally be constrained by [81]<br />

β<br />

β<br />

σ<br />

σ<br />

β<br />

β<br />

σ<br />

σ<br />

T − T


2.6 Transi<strong>en</strong>t diffusion and convection mass transport<br />

In this section we have applied the volume averaging method to solute transport with Soret<br />

effect in the case of a homog<strong>en</strong>eous medium in the β-phase. We now take the spatial<br />

average of ( 2-5), using the spatial averaging theorem [111] on the convective and diffusive<br />

terms. We begin our analysis with the <strong>de</strong>finition of two spatial <strong>de</strong>compositions for local<br />

conc<strong>en</strong>tration and velocity<br />

∂c<br />

∂t<br />

∂ c<br />

∂t<br />

β<br />

β<br />

+<br />

( c ) ( D c D T ) + ∇ ∇ = .<br />

∇.<br />

v ( 2-89)<br />

β<br />

β<br />

∇.<br />

D ∇c<br />

+ D ∇T<br />

β<br />

+ ∇.<br />

c v<br />

β<br />

β<br />

β<br />

Tβ<br />

+<br />

β<br />

∫<br />

n<br />

+<br />

∫<br />

c<br />

βσ β<br />

Aβσ<br />

n<br />

β<br />

β<br />

βσ<br />

Aβσ<br />

β<br />

v dA =<br />

Tβ ∇<br />

β<br />

( D ∇c<br />

+ D ∇T<br />

) dA<br />

β<br />

β<br />

Tβ<br />

51<br />

β<br />

( 2-90)<br />

∂ cβ<br />

+ ∇.<br />

cβ<br />

v β = ∇.<br />

Dβ<br />

∇cβ<br />

+ DTβ<br />

∇T<br />

( 2-91)<br />

β<br />

∂t<br />

In this equation we have consi<strong>de</strong>red the fluid flow to be incompressible, and have .ma<strong>de</strong><br />

use of no-slip boundary condition and equation ( 2-6) at the fluid-solid interface<br />

⎜<br />

⎛ β<br />

∂ ε ⎟<br />

⎞<br />

β cβ<br />

⎝ ⎠<br />

+ ∇.<br />

ε β<br />

∂t<br />

cβ<br />

v β<br />

⎛<br />

= ∇.<br />

⎜<br />

D ∇<br />

⎜<br />

⎜<br />

⎛<br />

β ε β cβ<br />

⎝<br />

⎝<br />

β<br />

⎟<br />

⎞<br />

D<br />

+<br />

⎠ V<br />

D<br />

⎞<br />

Tβ<br />

+<br />

⎟<br />

∫ n βσ Tβ<br />

dA<br />

V<br />

⎟<br />

Aβσ<br />

⎠<br />

β<br />

β<br />

A<br />

∫<br />

= ∇.<br />

D ∇c<br />

βσ<br />

n<br />

βσ<br />

c<br />

β<br />

β<br />

β<br />

dA + D<br />

+ D<br />

Tβ<br />

Tβ<br />

∇⎜<br />

⎛ε<br />

⎝<br />

∇T<br />

β<br />

β<br />

T<br />

β<br />

β<br />

We <strong>de</strong>fine the spatial <strong>de</strong>viations of the point conc<strong>en</strong>trations and velocities from the<br />

intrinsic phase average values by the relations<br />

c<br />

~<br />

β<br />

β = cβ<br />

+ cβ<br />

, in V β<br />

v<br />

β<br />

= v + ~ v , in V β<br />

β<br />

β<br />

β<br />

⎟<br />

⎞<br />

⎠<br />

( 2-92)<br />

( 2-93)<br />

The averaged quantities and their gradi<strong>en</strong>ts are tak<strong>en</strong> to be constants within the averaging<br />

volumeV , and this makes equation ( 2-92) to the form


∂<br />

c<br />

+ ε<br />

+ ε<br />

β<br />

β<br />

∂t<br />

−1<br />

β<br />

−1<br />

β<br />

+<br />

v<br />

β<br />

β .<br />

∇<br />

∇ε<br />

β . ⎜<br />

⎛ Dβ<br />

∇ c<br />

⎝<br />

c<br />

∇ε<br />

β . ⎜<br />

⎛ DTβ<br />

∇ T<br />

⎝<br />

β<br />

β<br />

β<br />

β<br />

β<br />

β<br />

= ∇.<br />

⎜<br />

⎛ Dβ<br />

∇ c<br />

⎝<br />

⎟<br />

⎞ + ∇.<br />

⎜<br />

⎛ DTβ<br />

∇ T<br />

⎠ ⎝<br />

⎟<br />

⎞ − ε<br />

⎠<br />

−1<br />

β<br />

β<br />

β<br />

β<br />

∇.<br />

c~<br />

~ v<br />

β<br />

β<br />

β<br />

⎟<br />

⎞ + ε<br />

⎠<br />

⎟<br />

⎞ + ε<br />

⎠<br />

52<br />

−1<br />

β<br />

−1<br />

β<br />

⎛ D<br />

∇.<br />

⎜<br />

⎜ V<br />

⎝<br />

β<br />

⎛ D<br />

∇.<br />

⎜<br />

⎜ V<br />

⎝<br />

Tβ<br />

A<br />

∫<br />

βσ<br />

A<br />

n<br />

∫<br />

βσ<br />

βσ<br />

n<br />

⎞<br />

c~<br />

β dA<br />

⎟<br />

⎟<br />

⎠<br />

βσ<br />

~<br />

⎞<br />

Tβ<br />

dA<br />

⎟<br />

⎟<br />

⎠<br />

( 2-94)<br />

Here, we have assumed that, as a first approximation, ~ v = 0 and c ~ = 0 . Therefore,<br />

the volume averaged convective transport has be<strong>en</strong> simplified to<br />

c = + ~c ~<br />

( 2-95)<br />

β β<br />

β v β ε β v β cβ<br />

β v β<br />

Subtracting Eq. ( 2-94) from Eq. ( 2-5) yields the governing equation for c β<br />

~<br />

∂c~<br />

β<br />

∂t<br />

+ ε<br />

+ ε<br />

+ v<br />

−1<br />

β<br />

−1<br />

β<br />

β<br />

. ∇c~<br />

⎛ D<br />

∇.<br />

⎜<br />

⎜ V<br />

⎝<br />

β<br />

+ ~ v . ∇ c<br />

β<br />

β<br />

β<br />

β<br />

β<br />

−1<br />

β<br />

~<br />

( D ∇c~<br />

) + ∇.<br />

( D ∇T<br />

)<br />

β<br />

β<br />

−1<br />

β −1<br />

Tβ<br />

∫ n βσ cβ<br />

dA + ε β ∇ε<br />

β . ⎜ Dβ<br />

cβ<br />

⎟ + ε β ∇.<br />

⎟ ⎝ ⎠ ⎜ V<br />

A<br />

A<br />

βσ<br />

~<br />

∇ε<br />

β . ⎜<br />

⎛ DTβ<br />

∇ T<br />

⎝<br />

⎞<br />

⎟<br />

⎠<br />

⎟<br />

⎞ − ε<br />

⎠<br />

= ∇.<br />

∇.<br />

c~<br />

~ v<br />

β<br />

β<br />

⎛ ∇<br />

β<br />

Tβ<br />

⎞<br />

β<br />

β<br />

⎛<br />

⎜ D<br />

⎝<br />

∫<br />

βσ<br />

n<br />

βσ<br />

β<br />

~<br />

⎞<br />

Tβ<br />

dA<br />

⎟<br />

⎟<br />

⎠<br />

( 2-96)<br />

One can also use the <strong>de</strong>finitions of the spatial <strong>de</strong>viations to obtain boundary conditions for<br />

c β<br />

~ from the boundary condition in Eq. ( 2-6). At the σ-β interface, we find that c β<br />

~ satisfies<br />

the relations<br />

~<br />

β<br />

β<br />

BC1: − βσ . ( Dβ∇c~ β + DTβ∇T<br />

β ) = nβσ<br />

. ( Dβ∇<br />

cβ<br />

+ DTβ∇<br />

Tβ<br />

)<br />

n , at A βσ<br />

( 2-97)<br />

and at the <strong>en</strong>trance and exit surface<br />

BC2: c ~<br />

β = f ( r,<br />

t)<br />

, at A βe<br />

( 2-98)<br />

The spatial <strong>de</strong>viations must satisfy the additional constraint that their average values be<br />

zero, in accordance with their <strong>de</strong>finition in Eq. ( 2-93).<br />

β<br />

c ~ = 0<br />

( 2-99)<br />

β<br />

The spatial <strong>de</strong>viation field is subject to simplifications allowed the <strong>de</strong>velopm<strong>en</strong>t of a<br />

relatively simple closure scheme to relate spatial <strong>de</strong>viations to average conc<strong>en</strong>tration.<br />

The non-local diffusion and thermal diffusion terms can be discar<strong>de</strong>d on the basis of


ε<br />

ε<br />

−1<br />

β<br />

−1<br />

β<br />

⎛<br />

⎜ Dβ<br />

∇. ⎜ ∫n<br />

V<br />

⎝ Aβσ βσ<br />

⎛<br />

⎜ DTβ<br />

∇. ⎜ ∫ n<br />

V<br />

⎝ Aβσ<br />

⎞<br />

c~<br />

dA⎟<br />

β ⎟<br />

⎠<br />

βσ<br />

~<br />

⎞<br />

T dA<br />

⎟<br />

β ⎟<br />

⎠<br />


~<br />

β<br />

cβ Cβ<br />

β Sβ<br />

= b . ∇ c + b . ∇ T<br />

( 2-107)<br />

in which b Cβ<br />

and b Sβ<br />

are referred to as the closure variables which are specified by the<br />

following boundary value problems. In <strong>de</strong>veloping these equations, we have collected all<br />

terms proportional to<br />

in the form<br />

β<br />

cβ<br />

β<br />

∇ and ∇ T , and writt<strong>en</strong> equation and the boundary conditions<br />

a∇<br />

cβ<br />

+ b∇<br />

T = 0<br />

( 2-108)<br />

where a and b are expressions containing the vector functions in the constitutive equation.<br />

In or<strong>de</strong>r to satisfy Eq. ( 2-108), we set each of the terms a and b, individually equal to zero,<br />

and this gives rise to the following equations. The solution of these problems is all subject<br />

to the constraint of equation ( 2-99). This means that the volume integrals of the vector<br />

fields must be zero.<br />

Problem IIa<br />

2<br />

v β . ∇ bCβ<br />

+ ~ v β = Dβ∇ bCβ<br />

( 2-109)<br />

BC: − n βσ . Dβ∇b Cβ = nβσ<br />

Dβ<br />

, at A βσ<br />

( 2-110)<br />

Periodicity: ( r ) = b ( r)<br />

bC β + l i Cβ<br />

, i=1,2,3 ( 2-111)<br />

β<br />

Averages: b = 0<br />

( 2-112)<br />

Problem IIb<br />

v<br />

Cβ<br />

. ∇ b = D ∇ b + D ∇ b<br />

( 2-113)<br />

2<br />

2<br />

β Sβ<br />

β Sβ<br />

Tβ<br />

Tβ<br />

BC: n . ( D β∇b<br />

β + D β∇b<br />

β ) = nβσ<br />

. D β<br />

− , at A βσ<br />

βσ S T T<br />

T<br />

Periodicity: ( r ) = b ( r)<br />

Averages: b = 0<br />

b S β + l i Sβ<br />

, i=1,2,3<br />

β<br />

Sβ<br />

54<br />

( 2-114)<br />

( 2-115)<br />

( 2-116)<br />

The closure problem can be solved also in a Chang’s unit cell shown in Fig. 2-4, in this<br />

case, we can replace the periodic boundary conditions for bSβ and b Cβ<br />

by a Dirichlet<br />

boundary condition. Therefore, the closure problem for pure diffusion and for Chang’s unit<br />

cell becomes


Problem I for Chang’s unit cell<br />

2<br />

∇ Tβ<br />

Fig. 2-4. Chang’s unit cell<br />

b = 0<br />

( 2-117)<br />

BC1: b Tβ<br />

= bTσ<br />

, at A βσ<br />

( 2-118)<br />

BC2: − . ∇b<br />

= −n<br />

. k ∇b<br />

+ n . ( k − k )<br />

n βσ k β Tβ<br />

βσ σ Tσ<br />

βσ β σ , at βσ<br />

σ Tσ<br />

2<br />

55<br />

A ( 2-119)<br />

0 = k ∇ b<br />

( 2-120)<br />

BC3: b = 0 , at 2 r r = ( 2-121)<br />

Tβ<br />

Problem IIa for Chang’s unit cell<br />

2<br />

∇ b Cβ<br />

= 0<br />

( 2-122)<br />

BC1: − n βσ . Dβ∇b Cβ = nβσ<br />

Dβ<br />

, at A βσ<br />

( 2-123)<br />

BC1: b = 0 , at 2 r r = ( 2-124)<br />

Cβ<br />

Problem IIb for Chang’s unit cell<br />

2<br />

2<br />

β ∇ b Sβ<br />

+ DTβ<br />

Tβ<br />

0 = D ∇ b<br />

( 2-125)<br />

BC1: n . ( D β∇b<br />

β + D β∇b<br />

β ) = nβσ<br />

. D β<br />

− , at A βσ<br />

βσ S T T<br />

T<br />

BC2: b = 0 , at 2 r r =<br />

Sβ<br />

( 2-126)<br />

( 2-127)


2.6.2 Closed form<br />

By substituting c β<br />

~ and T β<br />

~ from <strong>de</strong>composition equations into Eq. ( 2-94) and imposing<br />

the local equilibrium condition, Eq. ( 2-35), the closed form of the convection-double<br />

diffusion equation can be expressed by<br />

∂ ε c<br />

β<br />

∂t<br />

β<br />

β<br />

∇ + .<br />

( c ) ( c<br />

T T ) ∇ + ∇ ∇ =<br />

β β<br />

*<br />

β<br />

*<br />

ε β vβ β . ε βD<br />

β . β ε βD<br />

β .<br />

where the total dispersion and total thermal-dispersion t<strong>en</strong>sors are <strong>de</strong>fined by<br />

D<br />

*<br />

β<br />

56<br />

( 2-128)<br />

⎛<br />

1<br />

⎞<br />

β<br />

= D<br />

⎜<br />

β I + n βσ bCβ<br />

dA<br />

⎟<br />

− ~ v β b<br />

⎜<br />

Cβ<br />

V ∫ ⎟<br />

( 2-129)<br />

β ⎝ Aβσ<br />

⎠<br />

*<br />

D<br />

⎛<br />

⎜ 1<br />

⎞<br />

⎟<br />

⎛<br />

⎜ 1<br />

⎞<br />

⎟ ~<br />

⎝ βσ ⎠ ⎝<br />

βσ ⎠<br />

β<br />

Tβ<br />

= Dβ n βσ b Sβ<br />

dA + DTβ<br />

I + n βσ bTβ<br />

dA − v β b<br />

⎜<br />

Sβ<br />

V ∫ ⎟ ⎜<br />

β<br />

V ∫ ⎟<br />

( 2-130)<br />

A<br />

β A<br />

The area integral of the functions in this equations multiplied by the unit normal from one<br />

phase to another have be<strong>en</strong> <strong>de</strong>fined as tortuosity, which can be writt<strong>en</strong> for isotropic media<br />

I 1<br />

= I +<br />

τ<br />

∫<br />

Vβ A<br />

βσ<br />

n<br />

b<br />

βσ Cβ<br />

dA<br />

We can <strong>de</strong>fine an effective diffusion t<strong>en</strong>sor, in the isotropic case, according to<br />

( 2-131)<br />

Dβ<br />

I<br />

Deff<br />

= ( 2-132)<br />

τ<br />

The influ<strong>en</strong>ce of hydrodynamic dispersion appears in the volume integral of the function<br />

multiplied by the spatial <strong>de</strong>viation in the velocity<br />

D<br />

Hyd.<br />

= − ~ v<br />

( 2-133)<br />

β<br />

β b Sβ<br />

Therefore, the total dispersion t<strong>en</strong>sor appearing in Eq. ( 2-128) is the sum of the effective<br />

diffusion coeffici<strong>en</strong>t and the dispersion t<strong>en</strong>sor as<br />

*<br />

β = Deff<br />

DHyd.<br />

( 2-134)<br />

D +<br />

For a diffusive regime, the hydrodynamic t<strong>en</strong>sor will be zero. If we look at the effective<br />

diffusion and thermal diffusion coeffici<strong>en</strong>ts in equations ( 2-129) and ( 2-130), we can<br />

conclu<strong>de</strong> that the only condition that will produce the same tortuosity effect for diffusion<br />

and thermal diffusion mechanism is


1<br />

∫<br />

Vβ A<br />

βσ<br />

n<br />

βσ<br />

b<br />

Sβ<br />

dA = 0<br />

The numerical results in the next chapter will solve this problem.<br />

2.6.3 Non thermal equilibrium mo<strong>de</strong>l<br />

57<br />

( 2-135)<br />

Wh<strong>en</strong> the thermal equilibrium is not valid, the temperature <strong>de</strong>viations are writt<strong>en</strong> in terms<br />

of the gradi<strong>en</strong>t of the average temperature in two phases and we can write the<br />

conc<strong>en</strong>tration <strong>de</strong>viation as<br />

β<br />

β<br />

cβ Cβ<br />

. ∇ cβ<br />

+ b Sββ<br />

. ∇ Tβ<br />

+ Sβσ<br />

σ<br />

σ<br />

β<br />

β σ<br />

( Tβ<br />

− Tσ<br />

)<br />

~ = b b . ∇ T − r<br />

( 2-136)<br />

By substitution of this new conc<strong>en</strong>tration <strong>de</strong>viation and temperature <strong>de</strong>viation from<br />

equation ( 2-80) in β-phase into our quasi-steady closure problem for the spatial <strong>de</strong>viation<br />

conc<strong>en</strong>tration, equations ( 2-103)-( 2-106), we obtain following boundary value problems<br />

for diffusion and thermal diffusion closure variables.<br />

We note that the problem for b Cβ<br />

is the same as problem IIa for thermal equilibrium-one<br />

equation mo<strong>de</strong>l. Here are all the closure problem to <strong>de</strong>termine the closure variableb Cβ<br />

,<br />

b Sββ<br />

, Sβσ<br />

media<br />

b and r β to mo<strong>de</strong>l a macroscopic scale coupled heat and mass transfer in porous<br />

Problem IIIa<br />

2<br />

v β . ∇ b Cβ + ~ v β = Dβ ∇ b Cβ<br />

( 2-137)<br />

BC: − n βσ . Dβ ∇bCβ<br />

= n βσ Dβ<br />

, at A βσ<br />

( 2-138)<br />

Periodicity: ( r ) = b ( r)<br />

bC β + l i Cβ<br />

, i=1,2,3 ( 2-139)<br />

β<br />

Averages: b = 0<br />

( 2-140)<br />

Problem IIIb<br />

v<br />

Cβ<br />

2<br />

2<br />

β . b Sββ<br />

= Dβ ∇ b Sββ<br />

+ DTβ<br />

∇ bTββ<br />

∇ ( 2-141)<br />

BC: n βσ . ( D β ∇b<br />

Sββ<br />

+ DTβ<br />

∇bTββ<br />

) = n βσ . DTβ<br />

− , at A βσ<br />

( 2-142)<br />

Periodicity: ( r ) = b ( r)<br />

b S ββ + l i Sββ<br />

, i=1,2,3 ( 2-143)<br />

β<br />

Sββ<br />

Averages: b = 0<br />

( 2-144)


Problem IIIc<br />

v<br />

2<br />

2<br />

β . b Sβσ<br />

= Dβ ∇ b Sβσ<br />

+ DTβ<br />

∇ bTβσ<br />

∇ ( 2-145)<br />

BC: . ( D ∇b<br />

+ D ∇b<br />

) = 0<br />

− βσ β Sβσ<br />

Tβ<br />

Tβσ<br />

n , at A βσ<br />

( 2-146)<br />

Periodicity: ( r ) = b ( r)<br />

b S βσ + l i Sβσ<br />

, i=1,2,3 ( 2-147)<br />

β<br />

Averages: b = 0<br />

( 2-148)<br />

Problem IIId<br />

β<br />

Sβσ<br />

2<br />

rβ = Dβ<br />

∇ rβ<br />

+ DTβ<br />

2<br />

v . ∇<br />

∇ s<br />

( 2-149)<br />

BC: . ( ∇r<br />

+ D ∇s<br />

) = 0<br />

− βσ Dβ β Tβ<br />

β<br />

β<br />

n , at A βσ<br />

( 2-150)<br />

Periodicity: r ( i ) = r ( r)<br />

β<br />

β<br />

β<br />

r + l β , i=1,2,3 ( 2-151)<br />

Averages: r = 0<br />

( 2-152)<br />

By substituting c β<br />

~ and T β<br />

~ from the <strong>de</strong>composition giv<strong>en</strong> by equations ( 2-80) and ( 2-136)<br />

into Eq. ( 2-94), the closed form of the convection-double diffusion equation for the nonequilibrium<br />

two-equation temperature mo<strong>de</strong>l case can be expressed by<br />

∂ ε β<br />

∂<br />

∇.<br />

ε<br />

β<br />

β<br />

cβ<br />

β β<br />

β σ<br />

+ ∇.<br />

( ε β v β cβ<br />

) − ∇.<br />

( uCβ<br />

. ( Tβ<br />

− Tσ<br />

)<br />

t<br />

*<br />

β<br />

*<br />

β<br />

*<br />

σ<br />

( Dβ<br />

. ∇ cβ<br />

+ DTββ<br />

. ∇ Tβ<br />

+ DTβσ<br />

. ∇ Tσ<br />

)<br />

where the effective t<strong>en</strong>sors are <strong>de</strong>fined by<br />

D<br />

*<br />

β<br />

58<br />

=<br />

( 2-153)<br />

⎛ 1<br />

⎞<br />

β<br />

= D ⎜<br />

β I + n βσ bCβ<br />

dA⎟<br />

− ~ v βb<br />

Cβ<br />

⎜ V ∫ ( 2-154)<br />

⎟<br />

⎝ β Aβσ<br />

⎠<br />

*<br />

D<br />

⎛<br />

⎜ 1<br />

⎞<br />

⎟<br />

⎛<br />

⎜ 1<br />

⎞<br />

⎟ ~<br />

⎝ βσ ⎠ ⎝<br />

βσ ⎠<br />

β<br />

Tββ<br />

= Dβ n βσ b Sββ<br />

dA + DTβ<br />

I + n βσ bTββ<br />

dA − v β b<br />

⎜<br />

Sββ<br />

V ∫ ⎟ ⎜<br />

β<br />

V ∫ ⎟<br />

( 2-155)<br />

A<br />

β A<br />

*<br />

D<br />

⎛<br />

⎜ 1<br />

⎞<br />

⎟<br />

⎛<br />

⎜ +<br />

1<br />

⎞<br />

⎟ ~<br />

⎝ βσ ⎠ ⎝<br />

βσ ⎠<br />

u<br />

β<br />

Tβσ<br />

= Dβ n βσ b Sβσ<br />

dA + DTβ<br />

I n βσ b Tβσ<br />

dA − v β b Sβσ<br />

⎜V<br />

∫ ⎟ ⎜<br />

β<br />

V ∫<br />

( 2-156)<br />

⎟<br />

A<br />

β A<br />

⎛<br />

⎜<br />

1<br />

β<br />

Cβ<br />

= −Dβ<br />

n βσ rβ<br />

dA − DTβ<br />

n βσ sβ<br />

dA + v β r<br />

⎜<br />

β<br />

V ∫ ⎟ ⎜<br />

β<br />

V ∫ ⎟<br />

( 2-157)<br />

A<br />

β A<br />

⎝<br />

βσ<br />

⎞<br />

⎟<br />

⎠<br />

⎛<br />

⎜<br />

⎝<br />

1<br />

βσ<br />

⎞<br />

⎟<br />

⎠<br />

~


2.7 Results<br />

In or<strong>de</strong>r to illustrate the main features of the proposed multiple scale analysis, we have<br />

solved the dim<strong>en</strong>sionless form of the closure problems, for thermal equilibrium case (I, IIa,<br />

IIb), on a simple unit cell, to <strong>de</strong>termine the effective properties. We note here that the<br />

resolution of all mo<strong>de</strong>ls <strong>de</strong>scribed in the last section and comparison betwe<strong>en</strong> differ<strong>en</strong>t<br />

mo<strong>de</strong>ls are not the objective of this study.<br />

If we treat the repres<strong>en</strong>tative region as a unit cell in a spatially periodic porous medium, we<br />

can replace the boundary condition imposed at Aβe with a spatially periodic condition<br />

[111]. One such periodic porous media used in this study is shown in Fig. 2-5. The <strong>en</strong>tire<br />

phase system can be g<strong>en</strong>erated by translating the unit cell distances corresponding to the<br />

lattice base vectors l i , (i=1,2,3). The <strong>en</strong>tire set of equations can th<strong>en</strong> be solved within a<br />

single unit cell. The spatial periodicity boundary conditions are used in this study at the<br />

edges of the unit cell, as shown in Fig. 2-6.<br />

l β<br />

σ-phase<br />

59<br />

β-phase<br />

Fig. 2-5. Spatially periodic arrangem<strong>en</strong>t of the phases<br />

V<br />

V


Fig. 2-6. Repres<strong>en</strong>tative unit cell (εβ=0.8)<br />

This unit cell which will be used to compute the effective coeffici<strong>en</strong>ts is a symmetrical cell<br />

Fig. 2-6, for an or<strong>de</strong>red porous media (in line arrangem<strong>en</strong>t of circular cylin<strong>de</strong>rs). This type<br />

of geometry has already be<strong>en</strong> used for many similar problems [80, 111]. Th<strong>en</strong>, the macro-<br />

scale effective properties are <strong>de</strong>termined by equations ( 2-77), ( 2-129) and ( 2-130). For this<br />

illustration we have fixed the fluid mixture properties at ( ) ( ) = 1 ρc<br />

60<br />

ρc p p . The numerical<br />

simulations have be<strong>en</strong> done using the COMSOL TM Multiphysics finite elem<strong>en</strong>ts co<strong>de</strong>. In<br />

this study, we have calculated the longitudinal coeffici<strong>en</strong>ts which will be nee<strong>de</strong>d to<br />

simulate a test case for the macroscopic, one-dim<strong>en</strong>sional equation.<br />

2.7.1 Non-conductive solid-phase ( k ≈ 0 )<br />

σ<br />

In this section, the solid thermal conductivity is assumed to be very small and will be<br />

neglected in the equations. The corresponding closure problems and effective coeffici<strong>en</strong>ts<br />

in this case ( k ≈ 0 ) are listed as below<br />

σ<br />

Problem I ( k ≈ 0 ): closure problem for effective thermal conductivity coeffici<strong>en</strong>t<br />

σ<br />

2<br />

( ρcp ) vβ<br />

. bTβ<br />

+ ( ρcp<br />

) ~ vβ<br />

= kβ∇<br />

bTβ<br />

β<br />

∇ ( 2-158)<br />

β<br />

BC1: − n βσ . ∇bTβ<br />

= nβσ<br />

, at A βσ<br />

( 2-159)<br />

Periodicity: ( r ) = b ( r)<br />

b + l β , i=1,2,3 ( 2-160)<br />

T β i T<br />

β<br />

Averages: b = 0<br />

( 2-161)<br />

Tβ<br />

σ<br />

β


Problem IIa ( k ≈ 0 ): closure problem for the effective diffusion coeffici<strong>en</strong>t<br />

σ<br />

2<br />

vβ<br />

. ∇ bCβ<br />

+ ~ vβ<br />

= Dβ∇ bCβ<br />

( 2-162)<br />

BC: − n βσ . Dβ∇b Cβ = nβσ<br />

Dβ<br />

, at A βσ<br />

( 2-163)<br />

Periodicity: ( r ) = b ( r)<br />

bC β + l i Cβ<br />

, i=1,2,3 ( 2-164)<br />

β<br />

Averages: b = 0<br />

( 2-165)<br />

Cβ<br />

Problem IIb ( k ≈ 0 ): the closure problem for the effective thermal diffusion coeffici<strong>en</strong>t<br />

v<br />

σ<br />

. ∇ b = D ∇ b + D ∇ b<br />

( 2-166)<br />

2<br />

2<br />

β Sβ<br />

β Sβ<br />

Tβ<br />

Tβ<br />

BC: n βσ . ( D β∇b<br />

Sβ<br />

+ DTβ∇b<br />

Tβ<br />

) = nβσ<br />

. DTβ<br />

− , at A βσ<br />

( 2-167)<br />

Periodicity: ( r ) = b ( r)<br />

b + l β , i=1,2,3 ( 2-168)<br />

S β i S<br />

β<br />

Averages: b = 0<br />

( 2-169)<br />

Sβ<br />

and the effective coeffici<strong>en</strong>ts are calculated with<br />

k<br />

D<br />

⎛ 1<br />

⎞<br />

k ⎜ n βσ bTβ<br />

dA⎟<br />

− p b<br />

⎟<br />

β<br />

⎝<br />

βσ ⎠<br />

*<br />

β = β ε βI<br />

+<br />

⎜ V ∫<br />

A<br />

*<br />

β<br />

( ρc<br />

) ~ v β β<br />

T<br />

61<br />

( 2-170)<br />

⎛ 1<br />

⎞<br />

β<br />

= D ⎜<br />

β I + n βσ bCβ<br />

dA⎟<br />

− ~ v βb<br />

Cβ<br />

⎜ V ∫ ( 2-171)<br />

⎟<br />

⎝ β Aβσ<br />

⎠<br />

*<br />

D<br />

⎛<br />

⎜ 1<br />

⎞<br />

⎟<br />

⎛<br />

⎜ +<br />

1<br />

⎞<br />

⎟ ~<br />

⎝ βσ ⎠ ⎝<br />

βσ ⎠<br />

β<br />

Tβ<br />

= Dβ nβσb<br />

SβdA<br />

+ DTβ<br />

I nβσb<br />

TβdA<br />

− vβb<br />

Sβ<br />

⎜V<br />

∫ ⎟ ⎜<br />

β<br />

V ∫<br />

( 2-172)<br />

⎟<br />

A<br />

β A<br />

The macroscopic equations for<br />

β<br />

β<br />

T and<br />

β<br />

β<br />

c become<br />

β<br />

∂ ε T<br />

β<br />

β β<br />

( ) ( ) ⎜<br />

⎛ β β<br />

⎟<br />

⎞ = ∇ ⎜<br />

⎛ *<br />

ρc + ∇<br />

∇ ⎟<br />

⎞<br />

p<br />

ρc<br />

β<br />

p . ε<br />

β β β Tβ<br />

. ε β k . Tβ<br />

⎝<br />

⎠ ⎝<br />

⎠<br />

∂ ε c<br />

β<br />

∂t<br />

β<br />

β<br />

∂t<br />

∇ + .<br />

v ( 2-173)<br />

β β<br />

β<br />

β<br />

( ε β β cβ<br />

) ( ε β β cβ<br />

ε β Tβ Tβ<br />

) ∇ + ∇ ∇ =<br />

* *<br />

. D . D .<br />

v ( 2-174)<br />

One can find in the literature several expressions for the effective diffusion coeffici<strong>en</strong>t<br />

*<br />

base on the porosity, such as Wakao and Smith (1962) D β = ε β Dβ<br />

, Weissberg (1963)


D *<br />

β<br />

Dβ<br />

= , Maxwell (1881)<br />

1<br />

1− ln ε β<br />

2<br />

2D<br />

=<br />

3<br />

β<br />

D<br />

− ε β<br />

*<br />

β (see Quintard (1993)). For isotropic<br />

systems one may write β D<br />

*<br />

D as I τ , where τ is the scalar tortuosity of the porous<br />

matrix. The arbitrary, two dim<strong>en</strong>sional effective t<strong>en</strong>sor<br />

Φ<br />

*<br />

β<br />

* ⎡ϕ<br />

= ⎢<br />

*<br />

⎢ϕ<br />

⎣<br />

β<br />

xx<br />

β<br />

yx<br />

ϕ<br />

ϕ<br />

*<br />

β<br />

xy<br />

*<br />

β<br />

yy<br />

⎤<br />

⎥<br />

⎥<br />

⎦<br />

β<br />

62<br />

*<br />

Φ β is <strong>de</strong>fined as<br />

For the symmetric geometry shown in Fig. 2-6, wh<strong>en</strong> the Péclet number is zero, the<br />

effective coeffici<strong>en</strong>ts are also symmetric; therefore, we can write<br />

ϕ = ϕ = ϕ and<br />

* *<br />

ϕ = ϕ = 0.<br />

For a dispersive regime g<strong>en</strong>erated by a pressure gradi<strong>en</strong>t in the x-<br />

β<br />

xy<br />

β<br />

yx<br />

direction, the longitudinal dispersion coeffici<strong>en</strong>t ϕ is obviously more important than the<br />

transversal dispersion coeffici<strong>en</strong>t<br />

*<br />

β<br />

yy<br />

*<br />

β<br />

xx<br />

*<br />

β<br />

*<br />

β<br />

xx<br />

*<br />

β<br />

yy<br />

ϕ . In this study, as it is m<strong>en</strong>tioned in the previous<br />

section, we have just calculated the longitudinal coeffici<strong>en</strong>ts which will be nee<strong>de</strong>d to<br />

simulate a test case for the macroscopic, one-dim<strong>en</strong>sional equation.<br />

Fig. 2-7 shows our results of the closure problem resolution (A.I, A.IIa and A.IIb) in the<br />

case of pure diffusion ( Pe = 0 ).<br />

1<br />

τ<br />

*<br />

kβ<br />

εβkβ * *<br />

β DTβ<br />

=<br />

Dβ DTβ<br />

D<br />

=<br />

=<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

0.0 0.2 0.4 0.6 0.8 1.0<br />

ε (Porosity)<br />

Fig. 2-7. Effective diffusion, thermal diffusion and thermal conductivity coeffici<strong>en</strong>ts at Pe=0


We have therefore found that the effective thermal diffusion coeffici<strong>en</strong>t can also be<br />

estimated with this single tortuosity coeffici<strong>en</strong>t.<br />

* *<br />

Dβ<br />

DT<br />

=<br />

Dβ DT<br />

β<br />

β<br />

*<br />

k<br />

=<br />

ε k<br />

β<br />

β<br />

β<br />

I<br />

=<br />

τ<br />

63<br />

( 2-175)<br />

Here, the stared parameters are the effective coeffici<strong>en</strong>ts and the others are the coeffici<strong>en</strong>t<br />

in the free fluid.<br />

This relationship is similar to the one obtained for the effective diffusion and thermal<br />

conductivity in the literature [80, 111]. Therefore, we can say that the tortuosity factor acts<br />

in the same way on Fick diffusion coeffici<strong>en</strong>t and on thermal diffusion coeffici<strong>en</strong>t. In this<br />

case, the tortuosity is <strong>de</strong>fined as<br />

I 1<br />

1<br />

= I +<br />

T dA<br />

V ∫ nβσb<br />

β = I +<br />

τ<br />

V ∫ n<br />

β A<br />

β A<br />

βσ<br />

βσ<br />

b<br />

βσ Cβ<br />

dA<br />

( 2-176)<br />

This integral called tortuosity since, in the abs<strong>en</strong>ce of fluid flow, it modifies the diffusive<br />

properties of the system for the solute and heat transport.<br />

The results with convection ( Pe ≠ 0 ), are illustrated in Fig. 2-8. One can see that, for low<br />

Péclet number (diffusive regime), the ratio of effective diffusion coeffici<strong>en</strong>t to molecular<br />

diffusion coeffici<strong>en</strong>t in the porous medium is almost constant and equal to the inverse of<br />

the tortuosity of the porous matrix, which is consist<strong>en</strong>t with previously published results.<br />

On the opposite, for high Péclet numbers, the above m<strong>en</strong>tioned ratio changes following a<br />

power-law tr<strong>en</strong>d after a transitional regime. The curves of longitudinal mass dispersion<br />

(Fig. 2-8a) and thermal dispersion (Fig. 2-8b) have the classical form of dispersion curves<br />

[111]. In our case, the longitudinal mass and heat dispersion coeffici<strong>en</strong>ts can be<br />

repres<strong>en</strong>ted by<br />

*<br />

D β<br />

D<br />

β<br />

xx<br />

*<br />

k β<br />

=<br />

ε k<br />

β<br />

xx<br />

β<br />

=<br />

1<br />

1.<br />

20<br />

+<br />

0.<br />

0234<br />

Pe<br />

1.<br />

70<br />

( 2-177)<br />

where the dim<strong>en</strong>sionless Péclet number is <strong>de</strong>fined as<br />

Pe =<br />

β<br />

vβ<br />

lUC<br />

D<br />

( 2-178)<br />

β<br />

The dispersive part of the effective longitudinal thermal diffusion coeffici<strong>en</strong>t <strong>de</strong>creases<br />

with the Péclet number (Fig. 2-8c) and for high Péclet number it becomes negative. As we<br />

can see in Fig. 2-8c, there is a change of sign of the effective thermal diffusion coeffici<strong>en</strong>t.


This ph<strong>en</strong>om<strong>en</strong>on may be explained by the fact that, by increasing the fluid velocity, the<br />

gradi<strong>en</strong>t of<br />

b (x-coordinate of b Tβ<br />

) changes gradually its direction to the perp<strong>en</strong>dicular<br />

Tβ<br />

x<br />

flow path which could lead to a reversal the<br />

a result, a change of the<br />

*<br />

Tβ<br />

xx<br />

b (x-coordinate of b Sβ<br />

) distribution and as<br />

64<br />

Sβ<br />

x<br />

D sign (see Fig. 2-9).<br />

This curve can be fitted with a correlation as<br />

D<br />

*<br />

Tβ<br />

xx 1<br />

2.<br />

00<br />

= + 0.<br />

0052Pe<br />

( 2-179)<br />

DTβ 1.<br />

20<br />

The results in terms of Soret number, which is the ratio of isothermal diffusion coeffici<strong>en</strong>t<br />

on thermal diffusion coeffici<strong>en</strong>t, are original. Fig. 2-8d shows the ratio of effective Soret<br />

number to the Soret number in free fluid as a function of the Péclet number. The results<br />

show that, for a diffusive regime, one can use the same Soret number in porous media as<br />

the one in the free fluid ( S 1<br />

*<br />

T S T = ). This result agrees with the experim<strong>en</strong>tal results of<br />

xx<br />

Platt<strong>en</strong> and Costesèque (2004) and Costesèque et al. (2004) but, for convective regimes,<br />

the effective Soret number is not equal with the one in the free fluid. For this regime, the<br />

Soret ratio <strong>de</strong>creases with increasing the Péclet number, and for high Péclet number it<br />

becomes negative.<br />

To test the accuracy of the numerical solution, we have solved the steady-state vectorial<br />

closures A.I, A.IIa and A.IIb analytically for a plane Poiseuille flow betwe<strong>en</strong> two<br />

horizontal walls separated by a gap H.<br />

For this case, we found the following relation betwe<strong>en</strong> the effective longitudinal thermal<br />

diffusion and the thermal diffusion in the free fluid ( ε = 1)<br />

D<br />

D<br />

*<br />

Tβ<br />

xx<br />

Tβ<br />

2<br />

Pr Pe<br />

= 1−<br />

×<br />

Sc 210<br />

and the longitudinal dispersion is giv<strong>en</strong> (Wooding, 1960) by<br />

*<br />

D 2<br />

β<br />

xx Pe<br />

= 1+<br />

D 210<br />

β<br />

Here, Pe is <strong>de</strong>fined as<br />

β<br />

zβ<br />

v H<br />

Pe = where<br />

D<br />

β<br />

β<br />

β<br />

zβ<br />

( 2-180)<br />

( 2-181)<br />

v is the z-compon<strong>en</strong>t of the intrinsic<br />

average velocity of the fluid. The predicted values agree with the analytical results.


a<br />

b<br />

c<br />

d<br />

β<br />

D<br />

*<br />

β<br />

D<br />

β<br />

xx<br />

*<br />

β ε β k<br />

xx<br />

K<br />

*<br />

Tβ<br />

DTβ<br />

xx<br />

D<br />

T<br />

S<br />

*<br />

T<br />

S<br />

xx<br />

1.E+02<br />

1.E+01<br />

1.E+00<br />

1.E-01<br />

1.E+02<br />

1.E+01<br />

1.E+00<br />

1.E-01<br />

5<br />

-25<br />

-35<br />

-45<br />

-55<br />

1.E-02 1.E-01 1.E+00 1.E+01 1.E+02<br />

65<br />

Pe<br />

1.E-02 1.E-01 1.E+00 1.E+01 1.E+02<br />

Pe<br />

Diffusive regime<br />

-5<br />

1.E-02 1.E-01 1.E+00 1.E+01 1.E+02<br />

-15<br />

1.0<br />

0.6<br />

0.2<br />

-1.0<br />

Diffusive regime<br />

Pe<br />

Convective regime<br />

-0.2<br />

1.E-02<br />

-0.6<br />

1.E-01 1.E+00 1.E+01 1.E+02<br />

Pe<br />

Convective regime<br />

Fig. 2-8. Effective, longitudinal coeffici<strong>en</strong>ts as a function of Péclet number ( k ≈ 0 and ε = 0.<br />

8 ): (a) mass<br />

dispersion , (b) thermal dispersion , (c) thermal diffusion and (d) Soret number<br />

σ<br />

β


a) Pe=0.001, Arrow:<br />

b) Pe=10, Arrow:<br />

c) Pe=100, Arrow:<br />

b gradi<strong>en</strong>t Arrow:<br />

Tβ<br />

x<br />

Tβ<br />

x<br />

b gradi<strong>en</strong>t Arrow:<br />

b gradi<strong>en</strong>t Arrow:<br />

Tβ<br />

x<br />

66<br />

b Sβ<br />

x<br />

b Sβ<br />

b Sβ<br />

gradi<strong>en</strong>t<br />

x<br />

x<br />

gradi<strong>en</strong>t<br />

Fig. 2-9. Comparison of closure variables b and b for εβ=0.8<br />

Sβ<br />

x<br />

Tβ<br />

x<br />

gradi<strong>en</strong>t


2.7.2 Conductive solid-phase ( k ≠ 0 )<br />

σ<br />

In the previous section, we ma<strong>de</strong> the assumption k = 0 only for simplification whereas,<br />

for example, the soil thermal conductivity is about 0.52 W/m.K, and it <strong>de</strong>p<strong>en</strong>ds greatly on<br />

the solid thermal conductivity (in the or<strong>de</strong>r of 1 W/m.K) and varies with the soil texture.<br />

The thermal conductivity of most common non-metallic solid materials is about 0.05-20<br />

W/m.K, and this value is very large for metallic solids [47]. Values of k β for most<br />

common organic liquids range betwe<strong>en</strong> 0.10 and 0.17 W/m.K at temperatures below the<br />

normal boiling point, but water, ammonia, and other highly polar molecules have values<br />

several times as large [76].<br />

The increase of the effective thermal conductivity wh<strong>en</strong> increasing the phase conductivity<br />

ratio,κ , is well established from experim<strong>en</strong>tal measurem<strong>en</strong>ts and theoretical approaches<br />

([47, 111]) but the influ<strong>en</strong>ce of this ratio on thermal diffusion is yet unknown. In this<br />

section, we study the influ<strong>en</strong>ce of the conductivity ratio on the effective thermal diffusion<br />

coeffici<strong>en</strong>t. To achieve that, we solved numerically the closure problems with differ<strong>en</strong>t<br />

conductivity ratios. Fig. 2-10 shows the <strong>de</strong>p<strong>en</strong>d<strong>en</strong>ce of the effective t<strong>en</strong>sors with the<br />

conductivity ratio, for differ<strong>en</strong>t Péclet numbers. As shown in Fig. 2-10a the effective<br />

conductivity initially increases with an increase in κ and th<strong>en</strong> reaches an asymptote. As<br />

*<br />

the Péclet number increases, convection dominates and the effect of κ on k ε β k β is<br />

noticeably differ<strong>en</strong>t. The transition betwe<strong>en</strong> the high and low Péclet number regimes<br />

*<br />

occurs around Pe = 10 (see also [47]). For higher Péclet numbers (Pe > 10), k ε β k β is<br />

<strong>en</strong>hanced by lowering κ , as shown in Fig. 2-10 a for Pe = 14 . Our results for<br />

Tβ<br />

Tβ<br />

D<br />

*<br />

*<br />

D have a similar behaviour as ε β k β<br />

67<br />

σ<br />

k . Fig. 2-10 b shows the influ<strong>en</strong>ce of the<br />

conductivity ratio on the effective thermal diffusion coeffici<strong>en</strong>ts for differ<strong>en</strong>t Péclet<br />

numbers. One can see that increasing the solid thermal conductivity increases the value of<br />

the effective thermal diffusion coeffici<strong>en</strong>t for low Péclet numbers.<br />

On the contrary, for high Péclet numbers (Pe > 10) increasing the thermal conductivity<br />

ratio <strong>de</strong>creases the absolute value of the effective thermal diffusion coeffici<strong>en</strong>t. As shown<br />

in Fig. 2-10b, the thermal conductivity ratio has no influ<strong>en</strong>ce on the thermal diffusion<br />

coeffici<strong>en</strong>ts for the pure diffusion case ( Pe = 0 ). As we can see also in Fig. 2-11, both


a<br />

b<br />

closure variables fields b Sβ<br />

and b Tβ<br />

change with the thermal conductivity ratio but<br />

coupling results <strong>de</strong>fined by Eq. ( 2-172) , wh<strong>en</strong> velocity field is zero, are constant.<br />

β<br />

ε β k<br />

*<br />

K<br />

xx<br />

*<br />

Tβ<br />

DTβ<br />

xx<br />

D<br />

2.4<br />

2.0<br />

1.6<br />

1.2<br />

0.8<br />

0.4<br />

0.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

1.E-03 1.E-02 1.E-01 1.E+00 1.E+01 1.E+02<br />

σ β k k<br />

1.E-03 1.E-02 1.E-01 1.E+00 1.E+01 1.E+02<br />

σ β k k<br />

68<br />

Pe=0<br />

Pe=5<br />

Pe=8<br />

Pe=10<br />

Pe=0<br />

Pe=5<br />

Pe=8<br />

Pe=10<br />

Fig. 2-10. The influ<strong>en</strong>ce of conductivity ratio (κ ) on (a) effective, longitudinal thermal conductivity and (b)<br />

effective thermal diffusion coeffici<strong>en</strong>ts (εβ=0.8)<br />

Fig. 2-12 shows the closure variables fields<br />

b and<br />

Tβ<br />

x<br />

b Sβ<br />

x<br />

for a Péclet number equal to<br />

14, we can see also that both closure variables change with the thermal conductivity ratio.<br />

We can also solve the closure problem in a Chang’s unit cell (Fig. 2-4). In the closure<br />

problem we have a Dirichlet boundary condition in place of a periodic boundary. We have<br />

solved the closure problems giv<strong>en</strong> by Eqs. ( 2-117)-( 2-121) for the thermal conductivity<br />

coeffici<strong>en</strong>ts and Eqs. ( 2-125)-( 2-127) for the thermal diffusion coeffici<strong>en</strong>t for pure


diffusion. Fig. 2-13 shows the effective thermal diffusion and thermal conductivity versus<br />

the thermal conductivity ratio. One can see here also that, for pure diffusion, changing the<br />

conductivity ratio does not change the effective values.<br />

a) κ = 0.<br />

001 , b β<br />

b β<br />

T<br />

b) κ = 10 , b Tβ<br />

b Sβ<br />

c) κ = 100 , b Tβ<br />

b<br />

Sβ<br />

Fig. 2-11. Comparison of closure variables fields b Tβ<br />

and Sβ<br />

Pe 0 & ε = 0.<br />

8<br />

at pure diffusion ( )<br />

= β<br />

69<br />

S<br />

b for differ<strong>en</strong>t thermal conductivity ratio ( κ )


a) κ = 0.<br />

001 ,<br />

b) κ = 10 ,<br />

c) κ = 100 ,<br />

b Tβ<br />

x<br />

b Tβ<br />

b Tβ<br />

x<br />

x<br />

Fig. 2-12. Comparison of closure variables fields<br />

ratio ( )<br />

κ at convective regime ( Pe<br />

14 & ε = 0.<br />

8)<br />

= β<br />

70<br />

b and<br />

Tβ<br />

x<br />

Sβ<br />

x<br />

b Sβ<br />

b Sβ<br />

b Sβ<br />

x<br />

x<br />

x<br />

b for differ<strong>en</strong>t thermal conductivity


β<br />

ϕ<br />

ϕ *<br />

1.6<br />

1.4<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

1.E-03 1.E-02 1.E-01 1.E+00 1.E+01 1.E+02 1.E+03<br />

σ β k k<br />

71<br />

*<br />

k<br />

D<br />

k β<br />

D<br />

*<br />

Tβ<br />

Tβ<br />

Fig. 2-13. The influ<strong>en</strong>ce of conductivity ratio (κ ) on the effective coeffici<strong>en</strong>ts by resolution of the closure<br />

problem in a Chang’s unit cell (εβ=0.8 , Pe=0)<br />

2.7.3 Solid-solid contact effect<br />

It has be<strong>en</strong> emphasized in the literature that the “solid-solid contact effect” has a great<br />

consequ<strong>en</strong>ce on the effective thermal conductivity [79, 92], it can also change the effective<br />

thermal diffusion coeffici<strong>en</strong>t.<br />

In or<strong>de</strong>r to mo<strong>de</strong>l the effect of particle-particle contact we used the mo<strong>de</strong>l illustrated in Fig.<br />

2-14, in which the particle-particle contact area is <strong>de</strong>termined by the adjustable parameter<br />

a/d (the fraction of particle-particle contact area).<br />

Wh<strong>en</strong> a/d=0 th<strong>en</strong> the β-phase is continuous and the ratio β k<br />

*<br />

k becomes constant for<br />

large values of κ . Wh<strong>en</strong> a/d is not zero, at large values of κ , the solution predicts a linear<br />

<strong>de</strong>p<strong>en</strong>d<strong>en</strong>ce of β k<br />

*<br />

k on the ratio κ . The calculated results for both the continuous β-<br />

phase (non-touching particles) and the continuous σ-phase (touching particles) are shown<br />

in Fig. 2-15. The comparison pres<strong>en</strong>ted by Nozad et al. (1985) showed a very good<br />

agreem<strong>en</strong>t betwe<strong>en</strong> theory and experim<strong>en</strong>t. Sahraoui and Kaviany (1993) repeated the<br />

computation of Nozad et al. and find that the selection of a/d=0.002 gives closest<br />

agreem<strong>en</strong>t with experim<strong>en</strong>ts.


β<br />

k<br />

*<br />

k<br />

1.E+03<br />

1.E+02<br />

1.E+01<br />

1.E+00<br />

1.E-01<br />

1.E-02<br />

Fig. 2-14. Spatially periodic mo<strong>de</strong>l for solid-solid contact<br />

b) Touching particles<br />

geometry<br />

1.E-03<br />

1.E-05 1.E-04 1.E-03 1.E-02 1.E-01 1.E+00 1.E+01 1.E+02 1.E+03 1.E+04 1.E+05<br />

72<br />

σ β k k<br />

d<br />

C<br />

a) Non-touching<br />

particles geometry<br />

C/d=0.002<br />

Fig. 2-15. Effective thermal conductivity for (a) non-touching particles, a/d=0 (b) touching particles,<br />

a/d=0.002, (εβ=0.36, Pe=0)


Unfortunately we cannot use this type of geometry to study the effect of solid-solid<br />

connection on effective thermal diffusion coeffici<strong>en</strong>t because the fluid phase is not<br />

continuous. We have used therefore a geometry which has only particle connection in the<br />

x-direction as shown in Fig. 2-16.<br />

Fig. 2-16. Spatially periodic unit cell to solve the thermal diffusion closure problem with solid-solid<br />

connections a/d=0.002, (εβ=0.36, Pe=0)<br />

Fig. 2-17 shows the results for the effective coeffici<strong>en</strong>t obtained from the resolution of the<br />

closure problem for pure diffusion on the unit cell shown in Fig. 2-16. The a/d ratio has<br />

be<strong>en</strong> selected to be 0.002. It is clear from Fig. 2-17 that, while the particle connectivity<br />

changes greatly the effective values, the effective thermal diffusion coeffici<strong>en</strong>ts is<br />

in<strong>de</strong>p<strong>en</strong>d<strong>en</strong>t of the solid connectivity.<br />

73


β<br />

ϕ<br />

ϕ *<br />

1.E+03<br />

1.E+02<br />

1.E+01<br />

1.E+00<br />

1.E-01<br />

1.E-03 1.E-02 1.E-01 1.E+00 1.E+01 1.E+02 1.E+03<br />

σ β k k<br />

74<br />

D<br />

*<br />

k<br />

*<br />

Tβ<br />

Tβ<br />

Fig. 2-17. Effective thermal conductivity and thermal diffusion coeffici<strong>en</strong>t for touching particles, a/d=0.002,<br />

εβ=0.36, Pe=0<br />

The same problems have be<strong>en</strong> solved on the geometry shown in Fig. 2-14 but without y-<br />

connection parts.<br />

Comparison of closure variables fields b Tβ<br />

and b Sβ<br />

wh<strong>en</strong> the solid phase is continuous, for<br />

differ<strong>en</strong>t thermal conductivity ratios ( κ ) and pure diffusion are shown in Fig. 2-18. As we<br />

can see the closure variable for conc<strong>en</strong>tration b Sβ<br />

k β<br />

also change with conductivity ratio.<br />

The effective thermal conductivity and thermal diffusion coeffici<strong>en</strong>ts for this closure<br />

problem are plotted in Fig. 2-19, <strong>de</strong>spite the results illustrated in Fig. 2-15, the ratio<br />

*<br />

k k becomes constant for small values of κ because in y-direction there is not any<br />

β<br />

particle-particle resistance. At large values of κ the solution predicts a linear <strong>de</strong>p<strong>en</strong>d<strong>en</strong>ce<br />

of β k<br />

*<br />

k , on the ratio κ . However, the ratio Tβ<br />

Tβ<br />

D<br />

*<br />

D remains constant with the thermal<br />

conductivity ratio κ .<br />

D


a) κ = 0.<br />

001 , b Tβ<br />

b Sβ<br />

b) κ = 1,<br />

b Tβ<br />

b Sβ<br />

c) κ = 1000 , b Tβ<br />

b Sβ<br />

Fig. 2-18. Comparison of closure variables fields Tβ<br />

differ<strong>en</strong>t thermal conductivity ratio ( κ ) at pure diffusion<br />

b and b Sβ<br />

wh<strong>en</strong> the solid phase is continue, for<br />

75


β<br />

ϕ<br />

ϕ *<br />

1.E+04<br />

1.E+03<br />

1.E+02<br />

1.E+01<br />

1.E+00<br />

Touching particles<br />

in x -direction<br />

(C/d=0.002),<br />

d<br />

C<br />

&<br />

1.E-01<br />

1.E-05 1.E-04 1.E-03 1.E-02 1.E-01 1.E+00 1.E+01 1.E+02 1.E+03 1.E+04 1.E+05<br />

σ β k k<br />

76<br />

&<br />

D<br />

*<br />

k<br />

*<br />

Tβ<br />

Tβ<br />

Fig. 2-19. Effective thermal conductivity and thermal diffusion coeffici<strong>en</strong>t for touching particles, a/d=0.002,<br />

εβ=0.36<br />

2.8 Conclusion<br />

To summarize our findings, in this chapter we <strong>de</strong>termined the effective Darcy-scale<br />

coeffici<strong>en</strong>ts for heat and mass transfer in porous media using a volume averaging<br />

technique including thermal diffusion effects. We showed that the effective Soret number<br />

may <strong>de</strong>part from the micro-scale value because of advection effects. The results show that,<br />

for low Péclet numbers, the effective thermal diffusion coeffici<strong>en</strong>t is the same as the<br />

effective diffusion coeffici<strong>en</strong>t and that it does not <strong>de</strong>p<strong>en</strong>d on the conductivity ratio.<br />

However, in this regime, the effective thermal conductivity changes with the conductivity<br />

ratio. On the opposite, for high Péclet numbers, both the effective diffusion and thermal<br />

conductivity increase following a power-law tr<strong>en</strong>d, while the effective thermal diffusion<br />

coeffici<strong>en</strong>t <strong>de</strong>creases. In this regime, a change of the conductivity ratio will change the<br />

effective thermal diffusion coeffici<strong>en</strong>t as well as the effective thermal conductivity<br />

coeffici<strong>en</strong>t. At pure diffusion, ev<strong>en</strong> if the effective thermal conductivity <strong>de</strong>p<strong>en</strong>ds on the<br />

particle-particle contact, the effective thermal diffusion coeffici<strong>en</strong>t is always constant and<br />

in<strong>de</strong>p<strong>en</strong>d<strong>en</strong>t on the connectivity of the solid phase.<br />

k β<br />

D


Chapter 3<br />

Microscopic simulation and validation


3. Microscopic simulation and validation<br />

In this chapter, the macroscopic mo<strong>de</strong>l obtained by the theoretical method is validated by<br />

comparison with direct numerical simulations at the pore-scale. Th<strong>en</strong>, coupling betwe<strong>en</strong><br />

forced convection and Soret effect for differ<strong>en</strong>t cases is investigated.<br />

Nom<strong>en</strong>clature of Chapter 3<br />

A Area of the β-σ interface contained<br />

βσ<br />

within the macroscopic region, m 2<br />

A S<br />

c p<br />

segregation area, m 2<br />

78<br />

T β<br />

H<br />

β<br />

C<br />

Intrinsic average temperature in the βphase,<br />

K<br />

T , T Hot and cold temperature<br />

Constant pressure heat capacity, J.kg/K v Mass average velocity in the β-phase,<br />

β<br />

m/s<br />

c Total mass fraction in the β-phase x, y Cartesian coordinates, m<br />

β<br />

c β<br />

c 0<br />

β<br />

Intrinsic average mass fraction in the<br />

β-phase<br />

Initial conc<strong>en</strong>tration<br />

Greek symbols<br />

ε β<br />

Volume fraction of the β-phase or<br />

porosity<br />

Da Darcy number κ σ β k k , conductivity ratio<br />

Binary diffusion coeffici<strong>en</strong>t, m 2 /s<br />

Dynamic viscosity for the β-phase, Pa.s<br />

D β<br />

D Thermal diffusion coeffici<strong>en</strong>t, m<br />

Tβ<br />

2 /s.K<br />

μ β<br />

ρ β<br />

Total mass d<strong>en</strong>sity in the β-phase, kg/m 3<br />

*<br />

DT β<br />

Total thermal diffusion t<strong>en</strong>sor, m 2 /s.K τ Scalar tortuosity factor<br />

*<br />

D β<br />

Total dispersion t<strong>en</strong>sor, m 2 /s ψ Separation factor or dim<strong>en</strong>sionless Soret<br />

number<br />

k β<br />

k σ<br />

Thermal conductivity of the fluid<br />

phase, W/m.K<br />

Thermal conductivity of the solid<br />

phase, W/m.K<br />

Subscripts, superscripts and other symbols<br />

β<br />

Fluid-phase<br />

K β<br />

Permeability t<strong>en</strong>sor, m 2 * *<br />

k β , k<br />

σ Solid-phase<br />

Total thermal conductivity t<strong>en</strong>sors for<br />

no-conductive and conductive solid<br />

phase, W/m.K<br />

βσ β-σ interphase<br />

n βσ<br />

Unit normal vector directed from the βphase<br />

toward the σ –phase<br />

* Effective quantity<br />

Pe Cell Péclet number Spatial average<br />

S Soret number β<br />

T<br />

*<br />

S Effective Soret number<br />

T<br />

t Time, s<br />

T Temperature of the β-phase, K<br />

β<br />

Intrinsic β-phase average


3.1 Microscopic geometry and boundary conditions<br />

In or<strong>de</strong>r to validate the theory <strong>de</strong>veloped by the up-scaling technique in the previous<br />

chapter, we have compared the results obtained by the macro-scale equations with direct,<br />

pore scale, simulations. The porous medium is ma<strong>de</strong> of an array of the unit cell <strong>de</strong>scribed<br />

in Fig. 2-6. The array is chos<strong>en</strong> with 15 unit cells, as illustrated in Fig. 3-1.<br />

y<br />

x<br />

TH<br />

= 1<br />

Danckwerts B.C. for<br />

conc<strong>en</strong>tration field<br />

Fig. 3-1. Schematic of a spatially periodic porous medium ( T H : Hot Temperature and T C : Cold<br />

Temperature)<br />

In the macro-scale problem, the effective coeffici<strong>en</strong>ts are obtained from the previous<br />

solution of the closure problem. The macroscopic, effective coeffici<strong>en</strong>ts are the axial<br />

diagonal terms of the t<strong>en</strong>sor. Giv<strong>en</strong> the boundary and initial conditions, the resulting<br />

macro-scale problem is one-dim<strong>en</strong>sional.<br />

Calculations have be<strong>en</strong> carried out in the case of a binary fluid mixture with simple<br />

DTβ<br />

properties such that, ψ = ΔT<br />

× = 1<br />

D<br />

β<br />

and ( ρ p ) ( ρc<br />

) σ p β<br />

79<br />

c = .<br />

Microscopic scale simulations, as well as the resolution for the macroscopic problem, have<br />

be<strong>en</strong> performed using COMSOL Multiphysics TM .<br />

The 2D pore-scale dim<strong>en</strong>sionless equations and boundary conditions to be solved are Eqs.<br />

( 2-1)-( 2-9). Velocity was tak<strong>en</strong> to be equal to zero (no-slip) on every surface except at the<br />

<strong>en</strong>trance and exit boundaries. Danckwerts condition (Danckwerts, 1953) was imposed for<br />

the conc<strong>en</strong>tration at the <strong>en</strong>trance and exit (Fig. 3-1). In this dim<strong>en</strong>sionless system, we have<br />

imposed a thermal gradi<strong>en</strong>t equal to one.<br />

BC1: x = 0 n βe.<br />

( ∇cβ + ψ∇Tβ ) = 0 and T = TH<br />

= 1<br />

( 3-1)<br />

BC2: x = 15 n βe.<br />

( ∇cβ + ψ∇Tβ ) = 0 and T = TC<br />

= 0<br />

( 3-2)<br />

IC: t = 0 c = c 0 and T = T 0<br />

( 3-3)<br />

0 =<br />

0 =<br />

Initial condition ( T 0 = 0 & c 0 = 0 )<br />

TC<br />

= 0<br />

Danckwerts B.C. for<br />

conc<strong>en</strong>tration field


Mass fluxes are tak<strong>en</strong> equal to zero on other outsi<strong>de</strong> boundaries and on all fluid-solid<br />

boundary surfaces. Zero heat flux was used on the outsi<strong>de</strong> boundary except at the <strong>en</strong>trance<br />

and exit boundaries where we have imposed a thermal gradi<strong>en</strong>t. In the case of conductive<br />

solid-phase, the continuity boundary condition has be<strong>en</strong> imposed for heat flux on the fluid-<br />

solid boundary surface while these surfaces will be adiabatic for a no-conductive solid-<br />

phase case.<br />

Macroscopic fields are also obtained using the dim<strong>en</strong>sionless form of equations ( 2-17),<br />

( 2-18) ( 2-76) and ( 2-128).We obtained, from a method for predicting the permeability<br />

2<br />

t<strong>en</strong>sor [78], a Darcy number equal to Da = K β lUC<br />

= 0.<br />

25 , for the symmetric cell shown<br />

in Fig. 2-6.<br />

The boundary condition at the exit and <strong>en</strong>trance of the macro-scale domain were tak<strong>en</strong><br />

similar to the pore scale expressions but in terms of the averaged variables. Dep<strong>en</strong>ding on<br />

the pressure boundary condition and therefore the Péclet numbers, we can have differ<strong>en</strong>t<br />

flow regimes. First, we assume that the solid phase is not conductive (Section 3.2) and we<br />

compare the results of the theory with the direct simulation. Th<strong>en</strong>, the comparison will be<br />

done for a conductive solid-phase (Section 3.3) and differ<strong>en</strong>t Péclet numbers. In all cases,<br />

the micro-scale values are cell averages obtained from the micro-scale fields.<br />

3.2 Non-conductive solid-phase ( k ≈ 0 )<br />

σ<br />

In this section, the solid thermal conductivity is assumed to be very small and will be<br />

neglected in the equations and the solid-phase <strong>en</strong>ergy equation is not solved. Therefore we<br />

can express the equation of heat transfer Eq. ( 2-1) to Eq. ( 2-4) as<br />

∂T<br />

∂t<br />

β ( ρ ) + ( ρc<br />

) ∇ ( T ) = ∇.<br />

( k ∇T<br />

)<br />

c p β<br />

p<br />

BC1: . ( k T ) = 0<br />

βσ<br />

β ∇ β<br />

β<br />

. v , in the β-phase ( 3-4)<br />

β<br />

β<br />

β<br />

β<br />

n , at A βσ<br />

( 3-5)<br />

The various contributions of the fluid flow including pure diffusion and dispersion can be<br />

expressed as pres<strong>en</strong>ted in the followings.<br />

3.2.1 Pure diffusion ( Pe 0, k ≈ 0)<br />

≈ σ<br />

We have first investigated the Soret effect on mass transfer in the case of a static<br />

homog<strong>en</strong>eous mixture. In this case, we have imposed a temperature gradi<strong>en</strong>t equal to one,<br />

80


in the dim<strong>en</strong>sionless system, for the microscopic and macroscopic mo<strong>de</strong>ls, and we have<br />

imposed a Danckwerts boundary condition for conc<strong>en</strong>tration at the medium <strong>en</strong>trance and<br />

exit. The porosity ε β of the unit cell is equal to 0.8 and, therefore, in the case of pure<br />

diffusion, the effective coeffici<strong>en</strong>ts (diffusion, thermal conductivity and thermal diffusion)<br />

have be<strong>en</strong> calculated with a single tortuosity coeffici<strong>en</strong>t equal to 1.20 as obtained from the<br />

solution of the closure problem shown in Fig. 2-7.<br />

Fig. 3-2a shows the temporal evolution of the conc<strong>en</strong>tration at the exit for the two mo<strong>de</strong>ls,<br />

microscopic and macroscopic, with ( ψ = 1)<br />

and without ( ψ = 0 ) thermal diffusion.<br />

One can see that thermal diffusion modifies the local conc<strong>en</strong>tration and we cannot ignore<br />

this effect. The maximum modification at steady-state is equal to ψ . We also see that the<br />

theoretical results are here in excell<strong>en</strong>t agreem<strong>en</strong>t with the direct simulation numerical<br />

results.<br />

81


a<br />

c<br />

Mass fraction at the exit<br />

Volume averaged temperature.<br />

Volume averaged conc<strong>en</strong>tration.<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0.0<br />

-0.1<br />

0.9<br />

0.7<br />

0.5<br />

0.3<br />

0.1<br />

-0.1<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

-0.1<br />

-0.2<br />

-0.3<br />

-0.4<br />

-0.5<br />

ψ=1<br />

ψ=0<br />

0 20 40 60 80 100 120 140 160 180 200 220 240 260 280 300<br />

t=1<br />

t=0<br />

Time<br />

t=10 t=30<br />

82<br />

Prediction macro<br />

Averaged micro<br />

0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15<br />

Lines: prediction macro<br />

Points: averaged micro<br />

t=1<br />

t=0<br />

x<br />

t=10 t=30<br />

Lines: prediction macro<br />

Points: averaged micro<br />

t=300<br />

0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15<br />

x<br />

t=300<br />

Fig. 3-2. Comparison betwe<strong>en</strong> theoretical and numerical results at diffusive regime and κ=0, (a) time<br />

evolution of the conc<strong>en</strong>tration at x = 15 and (b and c) instantaneous temperature and conc<strong>en</strong>tration field


Fig. 3-2b and Fig. 3-2c show the distribution of temperature and conc<strong>en</strong>tration in the<br />

medium at giv<strong>en</strong> times. Here also, one can observe the change of the conc<strong>en</strong>tration profile<br />

g<strong>en</strong>erated by the Soret effect compared with the isothermal case (c=0), and the<br />

microscopic mo<strong>de</strong>l also perfectly fits the macroscopic results. These modifications are<br />

well matched with temperature profiles for each giv<strong>en</strong> time.<br />

3.2.2 Diffusion and convection ( Pe 0, k ≈ 0)<br />

≠ σ<br />

Next, we have imposed differ<strong>en</strong>t pressure gradi<strong>en</strong>ts on the system shown in Fig. 3-3. The<br />

temperature and conc<strong>en</strong>tration profiles at Pe = 1 for differ<strong>en</strong>t times are shown in Fig. 3-3a<br />

and Fig. 3-3b, respectively.<br />

The results show a significant change in the conc<strong>en</strong>tration profile because of species<br />

separation wh<strong>en</strong> imposing a thermal gradi<strong>en</strong>t. Here, also, the theoretical predictions are in<br />

very good agreem<strong>en</strong>t with the direct simulation of the micro-scale problem.<br />

Comparison of the conc<strong>en</strong>tration elution curves at x=0.5, 7.5 and 13.5 in Fig. 3-3c betwe<strong>en</strong><br />

the two regimes (with and without thermal diffusion) also shows that the elution curve for<br />

no-thermal diffusion is differ<strong>en</strong>t from the one with thermal diffusion. The shape of these<br />

curves is very differ<strong>en</strong>t from the pure diffusion case (Fig. 3-2a) because, in this case, the<br />

thermal diffusion process is changed by forced convection. One also observes a very good<br />

agreem<strong>en</strong>t betwe<strong>en</strong> the micro-scale simulations and the macro-scale predictions.<br />

83


a<br />

c<br />

Volume averaged conc<strong>en</strong>tration.<br />

Volume averaged temperature<br />

Volume averaged conc<strong>en</strong>tration.<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

-0.2<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

-0.2<br />

-0.4<br />

-0.6<br />

-0.8<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

-0.2<br />

-0.4<br />

-0.6<br />

-0.8<br />

0<br />

t=0<br />

t=1<br />

t=4<br />

0 2 4 6 8 10 12 14<br />

Lines: prediction macro<br />

Points: averaged micro<br />

t=1<br />

t=0<br />

x<br />

t=4<br />

84<br />

t=50<br />

t=10<br />

t=7<br />

t=16<br />

Lines: prediction macro<br />

Points: averaged micro<br />

t=13<br />

0 2 4 6 8 10 12 14<br />

x<br />

t=7<br />

0 5 10 15 20 25 30 35 40 45 50<br />

Time<br />

t=13<br />

t=16<br />

t=10<br />

t=50<br />

Macro (x=0.5)<br />

Macro (x=7.5)<br />

Macro (x=13.5)<br />

Micro (x=0.5)<br />

Micro (x=7.5)<br />

Micro (x=13.5)<br />

Isothermal<br />

Fig. 3-3. Comparison betwe<strong>en</strong> theoretical and numerical results, κ=0 and Pe=1, (a and b) instantaneous<br />

temperature and conc<strong>en</strong>tration field, (c) time evolution of the conc<strong>en</strong>tration at x = 0.5, 7.5 and 13.5


3.3 Conductive solid-phase ( k ≠ 0)<br />

σ<br />

Thermal properties of the solid matrix have also to be tak<strong>en</strong> into account in the thermal<br />

diffusion process. In this section, the heat diffusion through the solid-phase is consi<strong>de</strong>red.<br />

Therefore, the comparison has be<strong>en</strong> done for the same micro-scale mo<strong>de</strong>l but with a<br />

conductive solid-phase. First, we compare the results for a pure diffusion system and th<strong>en</strong><br />

we will <strong>de</strong>scribe the local dispersion coupling with Soret effect.<br />

3.3.1 Pure diffusion ( Pe 0, k ≠ 0)<br />

≈ σ<br />

Before we start to compare the theoretical results with the numerical one for the case of a<br />

conductive solid-phase, in or<strong>de</strong>r to see clearly the influ<strong>en</strong>ce of the thermal conductivity<br />

ratio on the separation process, we have solved the microscopic coupled heat and mass<br />

transport equations (Eqs. ( 2-1)-( 2-6)) in a simple geometry containing two unit cells. In the<br />

dim<strong>en</strong>sionless system shown in Fig. 3-4.<br />

Danckwerts conditions were imposed for the conc<strong>en</strong>tration at the <strong>en</strong>trance and exit (<br />

Fig. 3-1). In this dim<strong>en</strong>sionless system, we have imposed a horizontal thermal gradi<strong>en</strong>t<br />

equal to one.<br />

BC1: x = 0 n . ( c + ψ∇T ) = 0 and T = TH<br />

= 1<br />

( 3-6)<br />

βe<br />

∇ β<br />

β<br />

BC2: x = 2 n . ( c + ψ∇T ) = 0 and T = TC<br />

= 0<br />

( 3-7)<br />

βe<br />

∇ β<br />

β<br />

IC: t = 0 c = c 0 and T = T 0<br />

( 3-8)<br />

0 =<br />

0 =<br />

Mass fluxes are tak<strong>en</strong> equal to zero on other outsi<strong>de</strong> boundaries and on all fluid-solid<br />

boundary surfaces. The continuity boundary condition has be<strong>en</strong> imposed for heat flux on<br />

the fluid-solid boundary surface. Steady-state conc<strong>en</strong>tration and temperature fields for<br />

differ<strong>en</strong>t thermal conductivity ratio are repres<strong>en</strong>ted in Fig. 3-4. As we can see, the<br />

conc<strong>en</strong>tration distribution (or separation) <strong>de</strong>p<strong>en</strong>ds on the temperature distribution in the<br />

medium. Wh<strong>en</strong> the temperature distribution changes with the thermal conductivity, it will<br />

also change the conc<strong>en</strong>tration distribution.<br />

Fig. 3-5 shows the steady-state conc<strong>en</strong>tration and temperature profiles at y=0.5 (a section<br />

situated in the middle of the medium). We can see clearly that the thermal conductivity<br />

ratio changes the final temperature and conc<strong>en</strong>tration profiles. However, the final<br />

separations are constant. That means, although the thermal conductivity ratio change<br />

85


locally the conc<strong>en</strong>tration distribution in the medium, it has no influ<strong>en</strong>ce on the final<br />

separation. We discuss this point in more <strong>de</strong>tails below.<br />

a) κ = 0.<br />

001 , Temperature field Conc<strong>en</strong>tration field<br />

b) κ = 1,<br />

Temperature field Conc<strong>en</strong>tration field<br />

c) κ = 10 , Temperature field Conc<strong>en</strong>tration field<br />

Fig. 3-4. Influ<strong>en</strong>ce of the thermal conductivity ratio on the temperature and conc<strong>en</strong>tration fields<br />

86


Temperature<br />

Conc<strong>en</strong>tration<br />

1<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

-0.1<br />

-0.2<br />

-0.3<br />

-0.4<br />

-0.5<br />

0 0.5 1 1.5 2<br />

87<br />

x<br />

k=0.001<br />

k=1<br />

k=10<br />

k=0.001<br />

k=1<br />

k=10<br />

0 0.5 1 1.5 2<br />

Fig. 3-5. (a) Temperature and (b) conc<strong>en</strong>tration profiles for differ<strong>en</strong>t conductivity ratio<br />

x


The influ<strong>en</strong>ce of the thermal conductivity on the transi<strong>en</strong>t separation process is pres<strong>en</strong>ted<br />

in Fig. 3-6 for differ<strong>en</strong>t time steps.<br />

As we can see, the thermal conductivity of the solid phase locally changes the<br />

conc<strong>en</strong>tration profile. We must note here that one cannot judge from these results whether<br />

the thermal conductivity ratio has an influ<strong>en</strong>ce on the effective thermal diffusion<br />

coeffici<strong>en</strong>t or not. Actually, we must compare the average of the fields with the theoretical<br />

results as we will show in the following example.<br />

Conc<strong>en</strong>tration<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

-0.1<br />

-0.2<br />

-0.3<br />

-0.4<br />

-0.5<br />

0 0.5 1 1.5 2<br />

x<br />

88<br />

t=5<br />

t=0.5<br />

t=0.1<br />

C0=0<br />

Fig. 3-6. Temporal evolution of the separation profiles for differ<strong>en</strong>t thermal conductivity ratio<br />

In this example, the pure diffusion ( = 0)<br />

k=0.001<br />

k=1<br />

k=10<br />

Pe problem has be<strong>en</strong> solved for a ratio of<br />

conductivity equal to 10 ( κ = 10 ). In this condition, the local thermal equilibrium is valid<br />

as shown in Quintard et al. (1993). Th<strong>en</strong>, we can compare the results of the micro-scale<br />

mo<strong>de</strong>l and the macro-scale mo<strong>de</strong>l using only one effective thermal conductivity (local<br />

thermal equilibrium). We have shown in Section 3.3 that the thermal conductivity ratio has<br />

no influ<strong>en</strong>ce on the effective thermal diffusion coeffici<strong>en</strong>t for diffusive regimes. Therefore,<br />

we can use the same tortuosity factor for the effective diffusion and the thermal diffusion<br />

coeffici<strong>en</strong>t that the one used in the previous section ( β D D*<br />

*<br />

β = 0.<br />

83 and DTβ = 0.<br />

83DTβ<br />

).<br />

Whereas, we know that for pure diffusion, increasing the conductivity ratio increases the<br />

effective thermal conductivity. According to Fig. 2-10a for Pe = 0 and κ = 10,<br />

we obtain


a<br />

b<br />

k *<br />

= 1.<br />

72ε<br />

k . Fig. 3-7a and Fig. 3-7b show the temporal change in temperature and<br />

β<br />

β<br />

conc<strong>en</strong>tration profile for both mo<strong>de</strong>ls.<br />

Volume averaged temperature.<br />

Volume averaged conc<strong>en</strong>tration.<br />

0.9<br />

0.7<br />

0.5<br />

0.3<br />

0.1<br />

-0.1<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

-0.1<br />

-0.2<br />

-0.3<br />

-0.4<br />

-0.5<br />

t=1<br />

t=0<br />

t=10<br />

0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15<br />

Lines: prediction macro<br />

Points: averaged micro<br />

t=1<br />

t=0<br />

x<br />

t=30<br />

t=10 t=30<br />

89<br />

Lines: prediction macro<br />

Points: averaged micro<br />

t=300<br />

0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15<br />

x<br />

t=300<br />

Fig. 3-7. Comparison betwe<strong>en</strong> theoretical and numerical results at diffusive regime and κ=10, temporal<br />

evolution of (a) temperature and (b) conc<strong>en</strong>tration profiles


The symbols repres<strong>en</strong>t the direct numerical results (averages over each cell) and the lines<br />

are the results of the one-dim<strong>en</strong>sional macro-scale mo<strong>de</strong>l. One sees that, for a conductive<br />

solid-phase, our macro-scale predictions for conc<strong>en</strong>tration and temperature profiles are in<br />

excell<strong>en</strong>t agreem<strong>en</strong>t with the micro-scale simulations.<br />

In or<strong>de</strong>r to well un<strong>de</strong>rstand the effect of the thermal conductivity ratio on the thermal<br />

diffusion process, we have plotted in Fig. 3-8 the temperature and conc<strong>en</strong>tration profiles<br />

for differ<strong>en</strong>t thermal conductivity ratios, at a giv<strong>en</strong> time solution (t=10). As shown in this<br />

figure, a change in thermal conductivity changes the temperature profiles (Fig. 3-8a) and,<br />

consequ<strong>en</strong>tly, the conc<strong>en</strong>tration profiles (Fig. 3-8b). Since we showed that the thermal<br />

diffusion coeffici<strong>en</strong>t is constant at pure diffusion, we conclu<strong>de</strong> that modifications in<br />

conc<strong>en</strong>tration because of differ<strong>en</strong>t thermal conductivity ratio come from changing the<br />

temperature profiles. These modifications can be well distinguishable in Fig. 3-8c which<br />

shows the time evolution of the conc<strong>en</strong>tration at x = 15 for differ<strong>en</strong>t thermal conductivity<br />

ratio.<br />

90


c<br />

a<br />

Volume averaged temperature.<br />

Volume averaged conc<strong>en</strong>tration.<br />

Mass fraction at the exit<br />

0.9<br />

0.7<br />

0.5<br />

0.3<br />

0.1<br />

-0.1<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

-0.1<br />

-0.2<br />

-0.3<br />

-0.4<br />

-0.5<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0.0<br />

-0.1<br />

t=10<br />

0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15<br />

0 2 4 6 8 10 12 14<br />

ψ=1<br />

ψ=0<br />

x<br />

x<br />

t=10<br />

0 40 80 120 160 200 240 280<br />

Time<br />

91<br />

Prediction macro, κ=0<br />

Prediction macro, κ=1<br />

Prediction macro, κ=50<br />

Averaged micro, κ=0<br />

Averaged micro, κ=1<br />

Averaged micro, κ=50<br />

Prediction macro, κ=0<br />

Prediction macro, κ=1<br />

Prediction macro,κ=50<br />

Averaged micro, κ=0<br />

Averaged micro, κ=1<br />

Averaged micro, κ=50<br />

Prediction macro, κ=0<br />

Prediction macro, κ=1<br />

Prediction macro, κ=50<br />

Averaged micro, κ=0<br />

Averaged micro, κ=1<br />

Averaged micro, κ=50<br />

Fig. 3-8. Effect of thermal conductivity ratio at diffusive regime on (a and b) instantaneous temperature and<br />

conc<strong>en</strong>tration field at t=10 and (b) time evolution of the conc<strong>en</strong>tration at x = 15


3.3.2 Diffusion and convection ( Pe 0, k ≠ 0)<br />

≠ σ<br />

A comparison has be<strong>en</strong> ma<strong>de</strong> for Pe=1 and a thermal conductivity equal to 10. The<br />

temperature and conc<strong>en</strong>tration profiles for differ<strong>en</strong>t times are shown in Fig. 3-9a and Fig.<br />

3-9b, respectively. Here, also, the theoretical predictions are in very good agreem<strong>en</strong>t with<br />

the direct simulation of the micro-scale problem.<br />

Fig. 3-10a shows the effect of the Péclet number on the axial temperature distribution in<br />

the medium. At small Pe , the temperature distribution is linear, but as the pressure<br />

gradi<strong>en</strong>t (or Pe ) becomes large, convection dominates the axial heat flow.<br />

In Fig. 3-10b the steady-state distribution of the conc<strong>en</strong>tration is plotted for differ<strong>en</strong>t<br />

Péclet numbers. One can see clearly that the conc<strong>en</strong>tration profile changes with the Péclet<br />

number. For example, for Pe = 2 , because the medium has be<strong>en</strong> homog<strong>en</strong>ized thermally<br />

by advection in most of the porous domain, the conc<strong>en</strong>tration profile is almost the same as<br />

in the isothermal case (without thermal diffusion). Near the exit boundary, there is a<br />

temperature gradi<strong>en</strong>t which g<strong>en</strong>erates a consi<strong>de</strong>rable change in the conc<strong>en</strong>tration profile<br />

with an optimum point. This peak is a dynamic one resulting from coupling betwe<strong>en</strong><br />

convection and Soret effect.<br />

92


c<br />

a<br />

Volume averaged conc<strong>en</strong>tration.<br />

Volume averaged temperature<br />

Volume averaged conc<strong>en</strong>tration.<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

-0.2<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

-0.2<br />

-0.4<br />

-0.6<br />

-0.8<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

-0.2<br />

-0.4<br />

-0.6<br />

-0.8<br />

t=0<br />

t=1<br />

t=4<br />

0 2 4 6 8 10 12 14<br />

Lines: prediction macro<br />

Points: averaged micro<br />

t=1<br />

t=0<br />

t=4<br />

x<br />

93<br />

t=70<br />

t=10<br />

t=7<br />

t=16<br />

t=13<br />

Lines: prediction macro<br />

Points: averaged micro<br />

0 2 4 6 8 10 12 14<br />

x<br />

t=7<br />

0 5 10 15 20 25 30 35 40 45 50<br />

Time<br />

t=10<br />

t=13 t=16<br />

t=70<br />

Macro (x=0.5)<br />

Macro (x=7.5)<br />

Macro (x=13.5)<br />

Micro (x=0.5)<br />

Micro (x=7.5)<br />

Micro (x=13.5)<br />

Isothermal<br />

Fig. 3-9. Comparison betwe<strong>en</strong> theoretical and numerical results, κ=10 and Pe=1, (a) time evolution of the<br />

conc<strong>en</strong>tration at x = 0.5, 7.5 and 13.5 (b and c) instantaneous temperature and conc<strong>en</strong>tration field


a<br />

b<br />

Volume averaged conc<strong>en</strong>tration<br />

Volume averaged temperature<br />

0.5<br />

0.3<br />

0.1<br />

-0.1<br />

-0.3<br />

-0.5<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

0 2 4 6 8 10 12 14 16<br />

ψ=1<br />

0 2 4 6 8 10 12 14 16<br />

x<br />

x<br />

94<br />

Macro (pe=0.001)<br />

Macro (pe=0.1)<br />

Macro (pe=0.25)<br />

Macro (pe=0.75)<br />

Macro (pe=2)<br />

Micro (pe=0.001)<br />

Micro (pe=0.1)<br />

Micro (pe=0.25)<br />

Micro (pe=0.75)<br />

Micro (pe=2)<br />

Macro (pe=0.001)<br />

Macro (pe=0.1)<br />

Macro (pe=0.25)<br />

Macro (pe=0.75)<br />

Macro (pe=2)<br />

Micro (pe=0.001)<br />

Micro (pe=0.1)<br />

Micro (pe=0.25)<br />

Micro (pe=0.75)<br />

Micro (pe=2)<br />

Fig. 3-10. Influ<strong>en</strong>ce of Péclet number on steady-state (a) temperature and (b) conc<strong>en</strong>tration profiles (κ=10)


If we <strong>de</strong>fine a new parameter, A S , named segregation rate <strong>de</strong>fined as the surface betwe<strong>en</strong><br />

isothermal and thermal diffusion case conc<strong>en</strong>tration profiles, we can see that increasing the<br />

Péclet number <strong>de</strong>creases the segregation rate. The obtained puff conc<strong>en</strong>tration at the exit<br />

(x=15) and for differ<strong>en</strong>t Péclet numbers is illustrated in Fig. 3-11. The results show that<br />

the maximum separation passing trough the exit point increases with increasing the Péclet<br />

number while occuring in shorter period.<br />

Conc<strong>en</strong>tration at the exit<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

Pe =1.5<br />

Pe =0.75<br />

Pe =0.25<br />

0 20 40 60 80 100 120 140<br />

95<br />

Time<br />

Fig. 3-11. Influ<strong>en</strong>ce of Péclet number on steady-state conc<strong>en</strong>tration at the exit (κ=10)<br />

Pe =0<br />

As shown in Fig. 3-12 the conc<strong>en</strong>tration profile, and consequ<strong>en</strong>tly the peak point, not<br />

only <strong>de</strong>p<strong>en</strong>ds on the Péclet number, but also it is changed by the conductivity ratio, κ and<br />

separation factor, ψ .<br />

The conc<strong>en</strong>tration profile in the case of Pe = 0.<br />

75 and κ = 1 for differ<strong>en</strong>t separation<br />

factors, ψ , has be<strong>en</strong> plotted in Fig. 3-12a. One can see that increasing the separation<br />

factor increases the local segregation area of species. Fig. 3-12b shows the influ<strong>en</strong>ce of<br />

the conductivity ratio for a fixed Péclet number and separation factor ( Pe = 2 and ψ = 1 )<br />

on the conc<strong>en</strong>tration profile near the exit boundary ( x betwe<strong>en</strong> 10 and 15). The results<br />

show that a high conductivity ratio leads to smaller optimum point but higher segregation<br />

area than the i<strong>de</strong>al non-conductive solid-phase case. This means that the segregation area<br />

ψκ<br />

will be a function of . This specific result should be of importance in the analysis of<br />

Pe


a<br />

b<br />

species separation and especially in thermogravitational column, filled with a porous<br />

medium.<br />

Volume averaged conc<strong>en</strong>tartion.<br />

Volume averaged conc<strong>en</strong>tartion.<br />

0.2<br />

0<br />

-0.2<br />

-0.4<br />

-0.6<br />

-0.8<br />

0.2<br />

0.1<br />

0<br />

-0.1<br />

-0.2<br />

-0.3<br />

-0.4<br />

Pe=0.75<br />

κ=1<br />

0 3 6 9 12 15<br />

Pe=2<br />

ψ=1<br />

x<br />

10 11 12 13 14 15<br />

x<br />

96<br />

ψ=0<br />

ψ=1<br />

ψ=2<br />

ψ=3<br />

κ=0.1<br />

κ=1<br />

κ=10<br />

κ=100<br />

Fig. 3-12. Influ<strong>en</strong>ce of (a) separation factor and (b) conductivity ratio on pick point of the conc<strong>en</strong>tration<br />

profile


3.4 Conclusion<br />

In or<strong>de</strong>r to validate the theory <strong>de</strong>veloped by the up-scaling technique in the previous<br />

chapter, we have compared the results obtained by the macro-scale equations with direct<br />

pore-scale simulations. The porous medium is ma<strong>de</strong> of an array of unit cells. A good<br />

agreem<strong>en</strong>t has be<strong>en</strong> found betwe<strong>en</strong> macro-scale resolutions and micro-scale, direct<br />

simulations, which validates the proposed theoretical mo<strong>de</strong>l. We have pres<strong>en</strong>ted a situation<br />

illustrating how variations of Péclet number, conductivity ratio and separation factor<br />

coupled with Soret effect can change locally the segregation of species in a binary mixture.<br />

This may be of a great importance wh<strong>en</strong> evaluating the conc<strong>en</strong>tration in applications like<br />

reservoir <strong>en</strong>gineering, waste storage, and soil contamination.<br />

97


Chapter 4<br />

A new experim<strong>en</strong>tal setup to <strong>de</strong>termine<br />

the effective coeffici<strong>en</strong>ts


4. A new experim<strong>en</strong>tal setup to <strong>de</strong>termine the effective<br />

coeffici<strong>en</strong>ts<br />

The theoretical mo<strong>de</strong>l <strong>de</strong>veloped in chapter 2 concerning the effective thermal diffusion<br />

coeffici<strong>en</strong>t at pure diffusion regime confirmed that the tortuosity factor acts in the same<br />

way on both isothermal Fick diffusion coeffici<strong>en</strong>t and on thermal diffusion coeffici<strong>en</strong>t. We<br />

have shown also that the effective thermal diffusion coeffici<strong>en</strong>t does not <strong>de</strong>p<strong>en</strong>d on the<br />

solid to fluid conductivity ratio.<br />

In this study, a new experim<strong>en</strong>tal setup has be<strong>en</strong> <strong>de</strong>signed and fabricated to <strong>de</strong>termine<br />

directly the effective diffusion and thermal diffusion coeffici<strong>en</strong>ts for binary mixture. new<br />

experim<strong>en</strong>tal results obtained with a two-bulb apparatus are pres<strong>en</strong>ted. The diffusion and<br />

thermal diffusion of helium-nitrog<strong>en</strong> and helium-carbon dioxi<strong>de</strong> system through<br />

cylindrical samples filled with glass spheres of differ<strong>en</strong>t diameters and thermal<br />

conductivities are measured at the atmospheric pressure. Conc<strong>en</strong>trations are <strong>de</strong>termined by<br />

analysing the gas mixture composition in the bulbs with a katharometer <strong>de</strong>vice. A<br />

transi<strong>en</strong>t-state method for coupled evaluation of thermal diffusion and Fick coeffici<strong>en</strong>t in<br />

two bulbs system is proposed.<br />

99


Nom<strong>en</strong>clature of Chapter 4<br />

A<br />

A<br />

B<br />

*<br />

12<br />

*<br />

12<br />

*<br />

12<br />

0<br />

1<br />

,<br />

,<br />

C<br />

c , c<br />

0<br />

2<br />

Cross-sectional area of the connecting<br />

tube, m 2<br />

Ratios of collision integrals for<br />

calculating the transport coeffici<strong>en</strong>ts of<br />

mixtures for the L<strong>en</strong>nard-Jones (6-12)<br />

pot<strong>en</strong>tial<br />

Initial mass fraction of the heavier and<br />

lighter compon<strong>en</strong>t<br />

c Mass fraction of compon<strong>en</strong>t i in the b<br />

ib<br />

bulb<br />

c it<br />

∞<br />

c i<br />

o<br />

c<br />

Mass fraction of compon<strong>en</strong>t i in the t<br />

bulb<br />

Mass fraction of compon<strong>en</strong>t i at<br />

equilibrium<br />

Mass fraction at time t = 0<br />

100<br />

S Soret number, 1/K<br />

T<br />

D 12<br />

Diffusion coeffici<strong>en</strong>t, m 2 /s t<br />

*<br />

t<br />

Time, s<br />

d Diameter of the connecting tube, m<br />

*<br />

D<br />

T<br />

D<br />

*<br />

D T<br />

J i<br />

Effective diffusion coeffici<strong>en</strong>t, m 2 /s<br />

Thermal diffusion coeffici<strong>en</strong>t, m 2 /s<br />

Effective thermal diffusion coeffici<strong>en</strong>t,<br />

m 2 /s<br />

Mass diffusion flux, kg/m 2 .s<br />

k B<br />

Boltzmann constant, 1.38048 J/K<br />

k mix<br />

Thermal conductivity of the gas<br />

mixture, W.m/K<br />

k T<br />

*<br />

k T<br />

k<br />

l<br />

α<br />

M i<br />

m<br />

Thermal diffusion ratio<br />

Effective thermal diffusion ratio<br />

Thermal conductivity of the pure<br />

chemical species α , W.m/K<br />

L<strong>en</strong>gth of the connecting tube, m<br />

Molar mass of compon<strong>en</strong>t i, g/mol<br />

Particle shape factor<br />

*<br />

S T<br />

T<br />

T 0<br />

T ′<br />

*<br />

T<br />

T<br />

V<br />

V b<br />

Effective Soret number, 1/K<br />

Temperature of the col<strong>de</strong>r bulb, K<br />

Initial temperature, K<br />

Temperature of the hotter bulb, K<br />

Dim<strong>en</strong>sionless temperature<br />

Averaged temperature, K<br />

Diffusion relaxation time, s<br />

Volume of the bulb, m 3<br />

Volume of the bottom bulb, m 3<br />

V t Volume of the top bulb, m 3<br />

x α<br />

x<br />

β<br />

Greek symbols<br />

Mole fraction of species α<br />

Mole fraction of species β<br />

α Thermal diffusion factor<br />

T<br />

β Characteristic constant of the two-bulb<br />

diffusion cell <strong>de</strong>fined in Eq. ( 4-10), m -2<br />

Δ c Change in the conc<strong>en</strong>tration of heavier<br />

1<br />

compon<strong>en</strong>t at the steady state in the lower<br />

bulb<br />

ε Characteristic L<strong>en</strong>nard-Jones <strong>en</strong>ergy<br />

12<br />

parameter (maximum attractive <strong>en</strong>ergy<br />

betwe<strong>en</strong> two molecules), kg.m 2 /s 2<br />

ε<br />

Fractional void space (porosity)


N<br />

N i<br />

n<br />

The number of chemical species in the<br />

mixture<br />

Mass flux of compon<strong>en</strong>t i, kg/m 2 .s<br />

Number d<strong>en</strong>sity of molecules<br />

p Pressure, bar<br />

R Katharometer reading , mV<br />

K<br />

S Separation rate<br />

S, Q Quantities in the expression for α T<br />

S Partial gas saturation<br />

g<br />

μ α<br />

101<br />

Dynamic viscosity of pure species α ,<br />

g/cm.s<br />

σ Characteristic L<strong>en</strong>nard-Jones l<strong>en</strong>gth<br />

12<br />

(collision diameter), o<br />

A<br />

τ<br />

τ t<br />

Tortuosity<br />

Thermal diffusion relaxation time, s<br />

Φ The interaction parameter for gas-mixture<br />

αβ<br />

viscosity<br />

Ω<br />

Ω<br />

D<br />

( l,<br />

s)*<br />

Collision integral for diffusion<br />

Collision integral


4.1 Introduction<br />

In the previous chapters we <strong>de</strong>veloped a theoretical mo<strong>de</strong>l to predict effective thermal<br />

diffusion coeffici<strong>en</strong>ts from micro-scale parameters (thermal diffusion coeffici<strong>en</strong>t, porescale<br />

geometry, thermal conductivity ratio and Péclet numbers). The results confirm that<br />

for a pure diffusion regime, the effective Soret number in porous media is the same as the<br />

one in the free fluid [24, 25]. This means that the tortuosity factor acts in the same way on<br />

the Fick diffusion coeffici<strong>en</strong>t and on the thermal diffusion coeffici<strong>en</strong>t. In the pres<strong>en</strong>t work,<br />

the influ<strong>en</strong>ces of pore-scale geometry on effective thermal diffusion coeffici<strong>en</strong>ts in gas<br />

mixtures have be<strong>en</strong> measured experim<strong>en</strong>tally. Related to coupled-transport ph<strong>en</strong>om<strong>en</strong>a,<br />

the classical diffusion equation is completed with the additional thermal diffusion term.<br />

The mass flux, consi<strong>de</strong>ring a mono-dim<strong>en</strong>sional problem of diffusion, in the x -direction<br />

for a binary system, no subjected to external forces, and in which the pressure, but not the<br />

temperature, is uniform, can be writt<strong>en</strong><br />

⎡ ∂c1<br />

∂T<br />

⎤<br />

= −ρ<br />

⎢<br />

D12<br />

+ D<br />

⎣ ∂x<br />

∂x<br />

⎥<br />

( 4-1)<br />

⎦<br />

J1 β<br />

T<br />

where 12<br />

D is the ordinary diffusion coeffici<strong>en</strong>t and D T the thermal diffusion coeffici<strong>en</strong>t.<br />

Defining thermal diffusion ratio 12 D TD = , we can write (as in [52])<br />

J<br />

kT T<br />

⎡∂c1<br />

kT<br />

∂T<br />

⎤<br />

= −ρ<br />

D12<br />

⎢<br />

+<br />

⎣ ∂x<br />

T ∂x<br />

⎥<br />

( 4-2)<br />

⎦<br />

1 β<br />

Other quantities <strong>en</strong>countered are the thermal diffusion factor, α T , (for gases) and the Soret<br />

0 0<br />

coeffici<strong>en</strong>t, S T , <strong>de</strong>fined in literatures by α T = kT c1<br />

c2<br />

and ST = kT<br />

T respectively.<br />

Wh<strong>en</strong> k T in equation ( 4-2) is positive, heaviest species (1) moves toward the col<strong>de</strong>r<br />

region, and wh<strong>en</strong> it is negative, this species moves toward the warmer region. In some<br />

cases, there is a change in sign of the thermal diffusion ratio as the temperature is lowered<br />

(See [17] and [13]).<br />

By now, data for gas thermal diffusion in porous medium are not available and there is<br />

some uncertainty for the question concerning the relationship betwe<strong>en</strong> the effective liquid<br />

thermal diffusion coeffici<strong>en</strong>t and the micro-scale parameters (such as pore-scale geometry)<br />

[20, 74].<br />

In this study, using a gaseous mixture has the advantage that the relaxation time is much<br />

smaller compared to the one of liquid mixture.<br />

102


The main purpose of this part is to measure directly the binary diffusion and thermal<br />

diffusion coeffici<strong>en</strong>ts in porous media for the systems He-N2 and He-CO2, using a twobulb<br />

cell close to the <strong>de</strong>sign of Ney and Armistead [67]. This method has be<strong>en</strong> used<br />

already in many works to <strong>de</strong>termine transport properties in binary and ternary gases as<br />

well as liquids, with accurate results.<br />

4.2 Experim<strong>en</strong>tal setup<br />

In this study we have <strong>de</strong>signed and fabricated a new experim<strong>en</strong>tal setup that has be<strong>en</strong><br />

prov<strong>en</strong> suitable results for the study of diffusion and thermal diffusion in free fluid. It is an<br />

all-glass two-bulb apparatus, containing two double-spherical layers 1 (top) and 2 (bottom)<br />

as shown in Fig. 4-1. In fact, the particular differ<strong>en</strong>ce betwe<strong>en</strong> this system and the earliest<br />

two-bulb systems is that each bulb contains an interior glass sphere to serve as reservoir<br />

bulb and another exterior glass spheres in which, in the space betwe<strong>en</strong> two glass layers<br />

there may be a water circulation to regulate the reservoir temperature. As shown in Fig.<br />

4-2, the reservoir bulbs with equal and constant volume t = Vb<br />

= 1000<br />

103<br />

V cm 3 , joined by an<br />

insulated rigid glass tube of inner diameter d = 0.<br />

795cm<br />

and l<strong>en</strong>gth 8 cm containing a<br />

valve also ma<strong>de</strong> especially of 0.795 cm bore, and 5.87 cm long. Therefore, the total l<strong>en</strong>gth<br />

of the tube in which the diffusion processes occur is about l = 13.<br />

87 cm. To avoid<br />

convection, the apparatus was mounted vertically, with the hotter bulb uppermost.<br />

The conc<strong>en</strong>tration is <strong>de</strong>termined by analysing the gas mixture composition in each bulb<br />

with a katharometer. As we <strong>de</strong>scribed in Section 1.5.1, katharometer, or thermal<br />

conductivity <strong>de</strong>tector (Daynes 1933 [26], Jessop 1966 [46]), has already be<strong>en</strong> used to<br />

measure the conc<strong>en</strong>tration of binary gas mixtures. The method is based on the ability of<br />

gases to conduct heat and the property that the thermal conductivity of a gas mixture is a<br />

function of the conc<strong>en</strong>tration of its compon<strong>en</strong>ts. The thermal conductivity of a gas is<br />

inversely related to its molecular weight. Hydrog<strong>en</strong> has approximately six times the<br />

conductivity of nitrog<strong>en</strong> for example. The thermal conductivity of some gases with<br />

corresponding katharometer reading at atmospheric pressure is listed in Table 4-1.


Katharometer<br />

Manometer<br />

S<strong>en</strong>sor<br />

Diffusion zone<br />

Top<br />

Bulb<br />

Bottom<br />

Bulb<br />

Valve<br />

104<br />

Manometer<br />

Vacuum pump<br />

Bath temperature<br />

controller<br />

Vacuum pump<br />

He N 2<br />

Fig. 4-1. Sketch of the two-bulb experim<strong>en</strong>tal set-up used for the diffusion and thermal diffusion tests<br />

d=0.795 cm<br />

Vt= 1 liter<br />

Vb= 1 liter<br />

13.87 cm<br />

Fig. 4-2. Dim<strong>en</strong>sions of the <strong>de</strong>signed two-bulb apparatus used in this study


Table 4-1. Thermal conductivity and corresponding katharometer reading for some gases at atmospheric<br />

pressure and T=300°K<br />

Gas Air N2 CO2 He<br />

k(W/m.K) 0.0267 0.0260 0.0166 0.150<br />

RK(mV) 1122 1117 976 2345<br />

In this study, we have used the analyzer ARELCO-CATARC MP-R mo<strong>de</strong>l (Fig. 4-3) with<br />

a s<strong>en</strong>sor operating on the principle of thermal conductivity <strong>de</strong>tection. The electronics highperformance<br />

microprocessor of this <strong>de</strong>vice allows analysing the binary gas mixtures with<br />

±0.5% repeatability. The touch scre<strong>en</strong> display allows also seeing and verifying all ess<strong>en</strong>tial<br />

parameters e.g. scale analog output, temperature control, and access m<strong>en</strong>us. This type of<br />

Katharometer works with a circulation of the analyzed and refer<strong>en</strong>ce gases into the<br />

s<strong>en</strong>sors. The first series of the experim<strong>en</strong>t showed that the sampling with circulation<br />

cannot be applied in the two-bulb method because gas circulation can perturbs the<br />

establishm<strong>en</strong>t of the temperature gradi<strong>en</strong>t in the system. Small changes in the pressure in<br />

one bulb may produce forced convection in the system and cause a great error in the<br />

conc<strong>en</strong>tration evaluation. Therefore, in this study we have eliminated the pump system<br />

betwe<strong>en</strong> the bulbs and katharometer s<strong>en</strong>sors. Instead we connected the katharometer<br />

analyser s<strong>en</strong>sor directly to the bulbs as shown in Fig. 4-4. Therefore, the op<strong>en</strong> cell of the<br />

katharometer form a part of the diffusion cell, and so it can indicate continuously and<br />

without sampling the changes in composition as diffusion and thermal diffusion processes.<br />

The other s<strong>en</strong>sor of the katharometer has be<strong>en</strong> sealed perman<strong>en</strong>tly in air and the readings<br />

are the refer<strong>en</strong>ce readings.<br />

Fig. 4-3. Katharometer used in this study (CATARC MP – R)<br />

105


S<strong>en</strong>sor connections<br />

to Wheatstone<br />

bridge<br />

Refer<strong>en</strong>ce gas<br />

Fig. 4-4. A schematic of katharometer connection to the bulb<br />

For gases, the diffusion coeffici<strong>en</strong>t is inversely proportional to the absolute pressure and<br />

directly proportional to the absolute temperature to the 1.75 power as giv<strong>en</strong> by the Fuller et<br />

al. [33] correlation discussed in Reid et al. (1987).<br />

Pressure and temperature measurem<strong>en</strong>ts are ma<strong>de</strong> with two manometers and<br />

thermometers. The temperature of each bulb is kept at a constant value by circulating<br />

water from a bath temperature controller. In this study, for all diffusion measurem<strong>en</strong>ts, the<br />

temperature of two bulbs system is fixed to 300 °K. The gas purities are: He: 100%, N2:<br />

100% and CO2: 100%.<br />

4.2.1 Diffusion in a two-bulb cell<br />

The two-bulb diffusion cell is a simple <strong>de</strong>vice that can be used to measure diffusion<br />

coeffici<strong>en</strong>ts in binary gas mixtures. Fig. 4-5 shows a schematic of the two-bulb apparatus.<br />

V t<br />

Heated metal block<br />

Heated metal block<br />

A<br />

l<br />

Fig. 4-5. Two-bulb apparatus<br />

106<br />

V b<br />

Gas from the bulb to<br />

measure (analyzed gas)


Two vessels containing gases with differ<strong>en</strong>t compositions are connected by a capillary<br />

tube. The katharometer cell itself is connected with the bulb and its volume is negligible<br />

compared to the volume of the bulbs. The katharometer cell and the two bulbs were kept at<br />

a constant temperature of about 300 °C.<br />

The vacuum pumps are used at the beginning of the experim<strong>en</strong>t to eliminate the gas phase<br />

initially in the diffusion cell and in the gas flow lines.<br />

At the start of the experim<strong>en</strong>t (at t = 0), the valve is op<strong>en</strong>ed and the gases in the two bulbs<br />

can diffuse along the capillary tube. An analysis of binary diffusion in the two-bulb<br />

diffusion apparatus has be<strong>en</strong> pres<strong>en</strong>ted by Ney and Armistead (1947) [67] (see, also,<br />

Geankoplis, 1972). It is assumed that each bulb is at a uniform composition (the<br />

composition of each bulb is, of course, differ<strong>en</strong>t until equilibrium is reached). It is further<br />

assumed that the volume of the capillary tube connecting the bulbs is negligible in<br />

comparison to the volume of the bulbs themselves. This allows expressing the compon<strong>en</strong>t<br />

material balances for each bulb as follows<br />

dcib<br />

dcit<br />

ρ βVb<br />

= −ρ<br />

βVt<br />

= −N<br />

i A<br />

( 4-3)<br />

dt dt<br />

where A is the cross-sectional area of the capillary tube, c it is the mass fraction of<br />

compon<strong>en</strong>t i in the top bulb, and c ib is the mass fraction of that compon<strong>en</strong>t in the bottom<br />

bulb. The mass flux of species i through the capillary tube N i is consi<strong>de</strong>red to be positive<br />

if moving from top bulb to bottom bulb.<br />

The d<strong>en</strong>sity can be computed from the i<strong>de</strong>al gas law at the average temperature T<br />

P<br />

ρ β =<br />

( 4-4)<br />

RT<br />

at constant temperature and pressure the d<strong>en</strong>sity of an i<strong>de</strong>al gas is a constant; thus, there is<br />

no volume change on mixing and in the closed system the total flux N t must be zero.<br />

The composition in each bulb at any time is related to the composition at equilibrium ∞<br />

c i<br />

by<br />

( V + V ) c = V c + V c<br />

( 4-5)<br />

t<br />

b<br />

∞<br />

i<br />

t<br />

it<br />

b<br />

ib<br />

The compositions at the start of the experim<strong>en</strong>t are, therefore, related by<br />

∞<br />

0<br />

it<br />

( V t + Vb<br />

) ci<br />

= Vtc<br />

+ Vbcib<br />

0<br />

107<br />

( 4-6)


where<br />

0<br />

c is the mass fraction at time t = 0.<br />

In the analysis of Ney and Armistead it is assumed that, for i=1 at any instant, the flux 1 N<br />

is giv<strong>en</strong> by its one dim<strong>en</strong>sional, steady-state diffusion flux as<br />

ρ D<br />

( c − c )<br />

β 12<br />

J1 =<br />

1b<br />

1t<br />

l<br />

Thus ( 1 1 N J = ),<br />

108<br />

( 4-7)<br />

dc1b<br />

D12<br />

ρ βVb<br />

= −ρ<br />

β A(<br />

c1b<br />

− c1t<br />

)<br />

( 4-8)<br />

dt l<br />

To eliminate c1 t from Eq.( 4-8) one makes use of the compon<strong>en</strong>t material balance for<br />

both bulbs, Eqs. ( 4-6) and ( 4-7).<br />

dc1b<br />

∞<br />

= −β<br />

D12(<br />

c1b<br />

− c1<br />

)<br />

( 4-9)<br />

dt<br />

where β is a cell constant <strong>de</strong>fined by<br />

( V + V ) A<br />

=<br />

lVV<br />

t b β ( 4-10)<br />

t<br />

b<br />

A similar equation for the mass fraction of compon<strong>en</strong>t 2 in bulb t may also be <strong>de</strong>rived.<br />

Equation ( 4-9) is easily integrated, starting from the initial condition that at t = 0,<br />

to give<br />

0 ∞<br />

∞<br />

c1 = ( c1b<br />

− c1<br />

) exp( − D12t)<br />

+ c1<br />

o<br />

c1 b = c1b<br />

,<br />

b β ( 4-11)<br />

H<strong>en</strong>ce, if β is known th<strong>en</strong> just one value of c b is all that is nee<strong>de</strong>d to calculate the<br />

diffusivity D 12 . Alternatively, if an accurate value of 12<br />

D is available, Eq.( 4-11) can be<br />

used to calibrate a diffusion cell for later use in measuring diffusion coeffici<strong>en</strong>ts of other<br />

systems.<br />

In this study, the volume of the two bulbs is equal V t = Vb<br />

th<strong>en</strong>, we can write Eqs. ( 4-6) and<br />

( 4-10) as<br />

∞<br />

i<br />

0 0 ( c c ) 2<br />

c = +<br />

it<br />

ib<br />

( 4-12)<br />

2A<br />

β =<br />

( 4-13)<br />

lV<br />

where V is the bulb volume.


4.2.2 Two-bulb apparatus <strong>en</strong>d correction<br />

Wh<strong>en</strong> we <strong>de</strong>termine the diffusion coeffici<strong>en</strong>t in a two bulb system connected with a tube,<br />

the conc<strong>en</strong>tration gradi<strong>en</strong>t does not terminate at the <strong>en</strong>d of the connecting tube and,<br />

therefore an <strong>en</strong>d-correction has to be ma<strong>de</strong>. This correction was ma<strong>de</strong> in the calculation of<br />

the cell constants as an <strong>en</strong>d-effect by Ney and Armistead [67].They adjust the tube l<strong>en</strong>gth<br />

L for <strong>en</strong>d effects to give an effective l<strong>en</strong>gthl eff , giv<strong>en</strong> by<br />

l = l + 0.<br />

82d<br />

( 4-14)<br />

eff<br />

where d is the tube diameter.<br />

Rayleigh, 1945 [88], wh<strong>en</strong> investigating the velocity of sound in pipes, showed that one<br />

must add 0.82r for thick annulus flange and 0.52r for a thin annulus flange to each <strong>en</strong>d of<br />

the tube. Here, r is the tube radius.<br />

Wirz, 1947 showed that the <strong>en</strong>d corrections for sound in tubes <strong>de</strong>p<strong>en</strong>d on the annulus<br />

width, w, and diameter, d. The results fit the correlation [114]<br />

⎛ - 0.125d ⎞<br />

α = 0.<br />

60 + 0.<br />

22 exp⎜<br />

⎟<br />

( 4-15)<br />

⎝ w ⎠<br />

where α is the <strong>en</strong>d-correction factor.<br />

Analysis of many results on diffusion both in porous media and bulk gas also showed a<br />

significant differ<strong>en</strong>ce betwe<strong>en</strong> diffusion coeffici<strong>en</strong>ts measured in differ<strong>en</strong>t cells [108]. This<br />

differ<strong>en</strong>ce may arise through a differ<strong>en</strong>ce in geometry affecting the diffusion (say cell<br />

effect) or, in the case of the capillary tube, the <strong>en</strong>d correction factor being incorrect. More<br />

rec<strong>en</strong>t work indicates that the effect is due to differ<strong>en</strong>ces in cell geometry [106]. The<br />

exist<strong>en</strong>ce of this differ<strong>en</strong>ce implies that all measurem<strong>en</strong>ts of bulk gas diffusion by the two-<br />

bulb technique may contain systematic errors up to 2% [108].<br />

Arora et al, 1977 [4] using precise binary diffusion coeffici<strong>en</strong>ts showed that the <strong>en</strong>d<br />

correction formulation is not precise <strong>en</strong>ough wh<strong>en</strong> an accuracy of 0.1% in coeffici<strong>en</strong>ts is<br />

required. However, they proposed to calibrate the two-bulb cells with the standard<br />

diffusion coeffici<strong>en</strong>ts.<br />

According to this short bibliography, calculated diffusion coeffici<strong>en</strong>ts in a two-bulb<br />

apparatus <strong>de</strong>p<strong>en</strong>d on the cell geometry and <strong>en</strong>d connection tubes. Th<strong>en</strong>, in this study, we<br />

will use the standard values of diffusion coeffici<strong>en</strong>ts to calibrate the two-bulb apparatus for<br />

effective tube l<strong>en</strong>gth.<br />

109


In our work concerning the <strong>de</strong>termination of tortuosity, this error may be small because we<br />

have calculated a ratio of the two diffusion coeffici<strong>en</strong>ts. However, a better un<strong>de</strong>rstanding<br />

of this problem requires doing more experim<strong>en</strong>tal or numerical studies.<br />

4.2.3 Thermal diffusion in a two-bulb cell<br />

For calculation of the magnitu<strong>de</strong> of the Soret effect we used the same setup that we have<br />

used for diffusion processes. The diameter of the tube is small <strong>en</strong>ough to eliminate<br />

convection curr<strong>en</strong>ts and the volume of the tube is negligible in comparison with the<br />

volume of the bulbs.<br />

In the initial state, the whole setup is kept at a uniform and constant temperature T 0 and<br />

the composition of the mixture is uniform everywhere. After closing the valve in the tube,<br />

the temperature of the top bulb is increased to T H and the temperature of bottom bulb is<br />

lowered to T C , the two bulbs are set at the same pressure. After this intermediate state, the<br />

valve is op<strong>en</strong>ed. After a short time, a final stationary state is reached, in which there is a<br />

constant flux of heat from bulb t to bulb b. Measures have be<strong>en</strong> tak<strong>en</strong> such that T C and T H<br />

remain constant and, due to the Soret effect, it is observed a differ<strong>en</strong>ce in mass fraction<br />

betwe<strong>en</strong> the bulbs.<br />

Thermal diffusion separation is <strong>de</strong>termined by analysing the gas mixture composition in<br />

the bulbs by katharometric analysis.<br />

At steady-state, the separation due to thermal diffusion is balanced by the mixing effect of<br />

the ordinary diffusion, there is no net motion of either 1 or 2 species, so that J 0.<br />

If we<br />

take the tube axis to be in the x -direction, th<strong>en</strong> from Eq. ( 4-2) we get<br />

c kT<br />

∂T<br />

= −<br />

∂x<br />

T ∂x<br />

∂ 1<br />

110<br />

1 =<br />

( 4-16)<br />

We may ignore the effect of composition on k T and integrate this equation on temperature<br />

gradi<strong>en</strong>t betwe<strong>en</strong> T C and T H to get the change in conc<strong>en</strong>tration of the heavier compon<strong>en</strong>t<br />

at the steady state in the lower bulb [97].<br />

⎛ T ⎞ H<br />

Δc<br />

= − ⎜<br />

⎟<br />

1 kT<br />

ln<br />

( 4-17)<br />

⎝ TC<br />


th<strong>en</strong> the thermal diffusion factor α T is calculated from the following relation<br />

− Δc1<br />

α T =<br />

0 0 ⎛ T<br />

c c ln ⎜ 1 2<br />

⎝ T<br />

here, 0<br />

c 1 and<br />

H<br />

C<br />

⎞<br />

⎟<br />

⎠<br />

111<br />

( 4-18)<br />

0<br />

c 2 are the initial mass-fractions of the heavier and lighter compon<strong>en</strong>ts<br />

respectively in the binary gas mixture, and<br />

∞ ∞<br />

Δc1 = c1b<br />

− c1t<br />

.<br />

α T values thus obtained refer to an average temperature, T , in the range T C to T H ([37]<br />

and [95]) and these are <strong>de</strong>termined from the formula of Brown (1940) according to which<br />

[12]<br />

⎟ T ⎛ ⎞<br />

HTC<br />

TH<br />

T = ln ⎜<br />

( 4-19)<br />

TH<br />

− TC<br />

⎝ TC<br />

⎠<br />

which is based on an assumed temperature <strong>de</strong>p<strong>en</strong>d<strong>en</strong>ce for α T of the form<br />

The relaxation time τ t for this process can be expressed as [95]<br />

= a − bT<br />

−1<br />

α T<br />

.<br />

⎛ Vl<br />

⎞⎡<br />

TC<br />

⎤<br />

τ ⎜<br />

⎟<br />

t ≅ ⎢ ⎥ ( 4-20)<br />

⎝ D12<br />

A ⎠⎣TC<br />

+ TH<br />

⎦<br />

where V is the volume of one of the bulbs. The relaxation time is therefore proportional to<br />

the l<strong>en</strong>gth of the connecting tube, and inversely proportional to its cross-sectional area.<br />

The approach to the steady state is approximately expon<strong>en</strong>tial, and this was confirmed by<br />

following measurem<strong>en</strong>ts.<br />

The variation of pressure is small in each experi<strong>en</strong>ce. Theory and experim<strong>en</strong>t agree in<br />

showing that, at least at pressure below two atmospheres, the separation is in<strong>de</strong>p<strong>en</strong>d<strong>en</strong>t of<br />

the pressure; therefore in this study the thermal diffusion factor is not changed by small<br />

variation of pressure. In most gaseous mixture the thermal diffusion factor increases with<br />

increasing pressure. The temperature and conc<strong>en</strong>tration <strong>de</strong>p<strong>en</strong>d<strong>en</strong>ce of the thermal<br />

diffusion factor also were found to be affected by pressure [8].<br />

4.2.4 A transi<strong>en</strong>t-state method for thermal diffusion processes<br />

In this section, a transi<strong>en</strong>t-state method for thermal diffusion process in a two-bulb<br />

apparatus is proposed. In this case, the flux is the sum of Fick diffusion flux and thermal<br />

diffusion flux, as


( ) ⎟ D<br />

⎛ ⎞<br />

12<br />

D12kT<br />

TH<br />

J = − − ⎜<br />

1 ρ β c1t<br />

c1b<br />

ρβ<br />

ln<br />

( 4-21)<br />

l<br />

l ⎝ TC<br />

⎠<br />

at thermal equilibrium and for one-dim<strong>en</strong>sional case. Th<strong>en</strong> the conc<strong>en</strong>tration variation in<br />

the bottom bulb is giv<strong>en</strong> by<br />

( ) ⎟ dc ρ D<br />

D k ⎛ ⎞<br />

1b<br />

β 12 ρβ<br />

12 T TH<br />

ρ = − − + ⎜<br />

βVb<br />

c1b<br />

c1t<br />

ln<br />

( 4-22)<br />

dt l<br />

l ⎝ TC<br />

⎠<br />

The compositions at the starting of the experim<strong>en</strong>t are related by<br />

c = c<br />

( 4-23)<br />

0<br />

ib<br />

0<br />

it<br />

and, the composition in each bulb at any time is<br />

t<br />

it<br />

b<br />

ib<br />

0 ( V V ) c<br />

V c + V c = +<br />

( 4-24)<br />

t<br />

b<br />

ib<br />

Th<strong>en</strong> one can eliminate c1b from Eq. ( 4-22) using these two compon<strong>en</strong>t balances<br />

dc<br />

dt<br />

⎛<br />

⎞<br />

0 ⎜<br />

V ⎛ ⎞<br />

t TH<br />

+ β D =<br />

⎟<br />

⎜<br />

+ ⎜<br />

⎟<br />

12c1b<br />

D12<br />

c1b<br />

kT<br />

ln<br />

⎟<br />

( 4-25)<br />

⎝ Vt<br />

+ Vb<br />

⎝ TC<br />

⎠⎠<br />

1b β<br />

A similar equation for the mass fraction of compon<strong>en</strong>t 1 in bulb t may also be <strong>de</strong>rived. The<br />

integration of equation ( 4-25), starting from the initial condition at t = 0 ,<br />

c<br />

c<br />

⎡<br />

V<br />

k<br />

⎛ T<br />

⎜<br />

⎝<br />

⎞⎤<br />

⎟<br />

⎠⎦<br />

−D12βt<br />

( 1−<br />

e )<br />

112<br />

c = c , gives<br />

0<br />

1b ib<br />

0<br />

t<br />

H<br />

1b<br />

= 1b<br />

+ ⎢<br />

T ln ⎥<br />

Vt<br />

+ V ⎜<br />

b T ⎟<br />

( 4-26)<br />

C<br />

⎣<br />

If the value of D 12 is available, th<strong>en</strong> just one value of ( c1 b , t)<br />

is all that is nee<strong>de</strong>d to<br />

calculate the thermal diffusion factor and th<strong>en</strong> the thermal diffusion coeffici<strong>en</strong>t. However,<br />

wh<strong>en</strong> the experim<strong>en</strong>tal time evaluation of the conc<strong>en</strong>tration is available, both D 12 and<br />

k T (or T D ) can be evaluated. It is suffici<strong>en</strong>t to adjust 12 D and k T until equation ( 4-26) fits<br />

the experim<strong>en</strong>tal data.<br />

Wh<strong>en</strong> the volume of the two vessels is equal, Eq. ( 4-26) simplifies to<br />

t ⎛ − 2 ⎞<br />

0 S<br />

*<br />

= + ⎜ t<br />

c − ⎟<br />

1b<br />

c1b<br />

1 e<br />

2 ⎜ ⎟<br />

( 4-27)<br />

⎝ ⎠<br />

where, ⎟ ⎛ T ⎞ H S = k ⎜ T ln and t<br />

⎝ TC<br />

⎠<br />

lV<br />

AD<br />

diffusion relaxation time respectively.<br />

*<br />

= are a separation rate (or 1 c<br />

12<br />

Δ in Eq. ( 4-17)) and a


4.3 Experim<strong>en</strong>tal setup for porous media<br />

In a porous medium, the effective diffusion coeffici<strong>en</strong>t for solute transport is significantly<br />

lower than the free diffusion coeffici<strong>en</strong>t because of the constricted and tortuous solute flow<br />

paths. This effective diffusion coeffici<strong>en</strong>t is related to the free diffusion coeffici<strong>en</strong>t and<br />

tortuosity coeffici<strong>en</strong>t.<br />

The mono-dim<strong>en</strong>sional solute transport can be write as<br />

∂c<br />

= D<br />

∂t<br />

*<br />

2<br />

∂ c<br />

2<br />

∂x<br />

*<br />

D is the effective diffusion coeffici<strong>en</strong>t.<br />

113<br />

( 4-28)<br />

where<br />

We have used the same apparatus and method explained in the last section to measure the<br />

effective coeffici<strong>en</strong>ts except that, here, one part of the connecting tube (4 cm long,<br />

connected to bottom bulb) is filled with a synthetic porous medium ma<strong>de</strong> with the spheres<br />

of differ<strong>en</strong>t physical properties.<br />

4.4 Results<br />

4.4.1 Katharometer calibration<br />

To find the relative proportions of the compon<strong>en</strong>ts of a gas mixture, the instrum<strong>en</strong>t needs<br />

first to be calibrated. This is done by admitting mixtures of known proportions on the op<strong>en</strong><br />

cell and observing the differ<strong>en</strong>ce resistance betwe<strong>en</strong> refer<strong>en</strong>ce values and analyzed<br />

readings. The precision with which the change in composition of a mixture can be<br />

measured <strong>de</strong>p<strong>en</strong>ds, of course, on the differ<strong>en</strong>ce of the thermal conductivities of the two<br />

compon<strong>en</strong>ts and this also <strong>de</strong>p<strong>en</strong>ds on the differ<strong>en</strong>ce of the molecular masses. Fig. 4-6<br />

shows an example of katharometer calibration curve for mixtures of He − CO2<br />

, which<br />

have be<strong>en</strong> obtained in or<strong>de</strong>r to interpolate the changes in conc<strong>en</strong>tration as a function of<br />

katharometer readings. One can see that the katharometer calibration curve gives a very<br />

close approximation in shape to the theoretical curve of thermal conductivity against<br />

conc<strong>en</strong>tration.


Thermal conductivity of the mixture<br />

(W/m.K)<br />

0.138<br />

0.118<br />

0.098<br />

0.078<br />

0.058<br />

0.038<br />

0.018<br />

-145.4<br />

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1<br />

Mole fraction of CO 2<br />

Estimation<br />

Calibrations<br />

114<br />

1654.6<br />

1454.6<br />

1254.6<br />

1054.6<br />

Fig. 4-6. Katharometer calibration curve with related estimation of thermal conductivity values for the<br />

system He-CO2<br />

The thermal conductivities for gas mixtures at low d<strong>en</strong>sity have be<strong>en</strong> estimated by Mason-<br />

Sax<strong>en</strong>a approach [58]<br />

k<br />

mix<br />

=<br />

N<br />

∑<br />

α = 1∑<br />

854.6<br />

654.6<br />

454.6<br />

254.6<br />

xα<br />

kα<br />

( 4-29)<br />

x Φ<br />

β<br />

β<br />

αβ<br />

Here, the dim<strong>en</strong>sionless quantities Φ αβ are<br />

Φ<br />

αβ<br />

=<br />

1 ⎛<br />

⎜<br />

M<br />

1+<br />

8 ⎜<br />

⎝ M<br />

α<br />

β<br />

⎞<br />

⎟<br />

⎠<br />

−1<br />

2<br />

⎡ ⎛ ⎞<br />

⎢1<br />

⎜<br />

μα<br />

+ ⎟<br />

⎢ ⎜ ⎟<br />

⎣ ⎝ μ β ⎠<br />

1 2<br />

⎛ M<br />

⎜<br />

⎝ M<br />

1 4<br />

β<br />

α<br />

⎞<br />

⎟<br />

⎠<br />

⎤<br />

⎥<br />

⎥<br />

⎦<br />

2<br />

54.6<br />

Katharometer differ<strong>en</strong>ce reading (mV)<br />

( 4-30)<br />

where N is the number of chemical species in the mixture. For each species α, x α is the<br />

mole fraction, k α is the thermal conductivity, μ α is the viscosity at the system temperature<br />

and pressure, and M α is the molecular weight of species α.<br />

The properties of N2, CO2 and He required to calculate thermal conductivity of mixture<br />

have be<strong>en</strong> listed in Table 4-2 at 300°K and 1 atm.


Table 4-2. The properties of CO2, N2 and He required to calculate kmix<br />

M α<br />

4<br />

μ α × 10<br />

(g/cm.s)<br />

CO2 44.010 1.52 433<br />

N2 28.016 1.76 638<br />

115<br />

7<br />

k α × 10<br />

He 4.002 2.01 3561<br />

4.4.2 Diffusion coeffici<strong>en</strong>t<br />

cal/cm.s.K<br />

(T=300 °C, P=1 atm.)<br />

Usually, five mo<strong>de</strong>s of gas transport can be consi<strong>de</strong>red in porous media [59]. As illustrated<br />

schematically in Fig. 4-7, four of them are related to conc<strong>en</strong>tration, temperature or partial<br />

pressure gradi<strong>en</strong>ts (molecular diffusion, thermal diffusion, Knuds<strong>en</strong> diffusion and surface<br />

diffusion), and one to the total gas pressure gradi<strong>en</strong>t (viscous or bulk flow). Wh<strong>en</strong> the gas<br />

molecular mean free path becomes of the same or<strong>de</strong>r as the tube dim<strong>en</strong>sions, freemolecule,<br />

or Knuds<strong>en</strong>, diffusion becomes important. Due to the influ<strong>en</strong>ce of walls,<br />

Knuds<strong>en</strong> diffusion and configurational diffusion implicitly inclu<strong>de</strong> the effect of the porous<br />

medium.<br />

Fig. 4-7. Solute transport process in porous media<br />

In the discussion which follows, no total pressure gradi<strong>en</strong>t (no bulk flow) is consi<strong>de</strong>red<br />

since this is the condition which prevails in the experim<strong>en</strong>ts pres<strong>en</strong>ted in this study. In<br />

most of the former studies, surface diffusion was either neglected or consi<strong>de</strong>red only as a<br />

rapid process since its contribution to the overall transport cannot be assessed precisely.<br />

Knuds<strong>en</strong> diffusion is neglected because the pore size is larger than the l<strong>en</strong>gth of the free<br />

path of the gas molecules. For example, in the atmospheric pressure, the mean free path of


the helium molecule at 300 °C is about 1.39×10 -7 m. Thus, in this study, only binary<br />

molecular gas diffusion is consi<strong>de</strong>red.<br />

In this type of experim<strong>en</strong>t, it is assumed that the diffusion coeffici<strong>en</strong>t of the gas mixture is<br />

in<strong>de</strong>p<strong>en</strong>d<strong>en</strong>t of composition, and the transi<strong>en</strong>t temperature rises due to Dufour effects are<br />

insignificant. It is also assumed that the conc<strong>en</strong>tration gradi<strong>en</strong>t is limited to the connecting<br />

tube whereas the composition within each bulb remains uniform at all times. In addition,<br />

the pressure is assumed to be uniform throughout the cell, so that viscous effects are<br />

negligible, and high <strong>en</strong>ough to minimize free-molecular (Knuds<strong>en</strong>) diffusion.<br />

−5<br />

For the setup <strong>de</strong>scribed in Section 2, the cell constant β is equal to 7.<br />

16×<br />

10 , therefore we<br />

can rewrite Eq.( 4-11) as<br />

0 ∞<br />

−5<br />

∞<br />

c1 = ( c1b<br />

− c1<br />

) exp( −7.<br />

16×<br />

10 D12t)<br />

+ c1<br />

b ( 4-31)<br />

Using Eq. ( 4-31), only one point ( c t, t)<br />

1 is suffici<strong>en</strong>t to <strong>de</strong>termine the diffusion coeffici<strong>en</strong>t.<br />

The katharometer interval registration data has be<strong>en</strong> set to one minute; therefore, there is<br />

suffici<strong>en</strong>t data to fit Eq. ( 4-31) on the experim<strong>en</strong>tal data to obtain a more accurate<br />

coeffici<strong>en</strong>t, compared with a one point calculation. Th<strong>en</strong>, the obtained binary diffusion<br />

coeffici<strong>en</strong>t is about 0.690 cm 2 /s for 2 N He −<br />

116<br />

system and 0.611 cm2 /s for He − CO2<br />

system. In the literature [103], binary diffusion coeffici<strong>en</strong>t for a 2 N He − system measured<br />

with two-bulb method at the condition of p=101.325 kPa, and T=299.19 °K, is about<br />

0.7033 cm 2 /s. This coeffici<strong>en</strong>t for a He − CO2<br />

system has be<strong>en</strong> reported as 0.615 cm 2 /s at<br />

300°K [28]. Using these standard coeffici<strong>en</strong>ts a new calibrated mean cell constant has be<strong>en</strong><br />

calculated. This constant that will be used for all next experim<strong>en</strong>ts is equal to β/1.015.<br />

The theoretical estimation of the diffusion coeffici<strong>en</strong>ts also are not differ<strong>en</strong>t from values<br />

obtained in this study which show the validity of the measuring method and apparatus (the<br />

theoretical formulation has be<strong>en</strong> explained in App<strong>en</strong>dix A).<br />

Table 4-3 shows the necessary data to estimate the diffusion coeffici<strong>en</strong>t for the system,<br />

He − CO2<br />

and 2 N<br />

He − . The calculation of mixture parameters, dim<strong>en</strong>sionless<br />

temperature, collision integral and diffusion coeffici<strong>en</strong>t from Eq. (A. 2) and for<br />

temperatures applied in this study have be<strong>en</strong> listed in Table 4-4.


Table 4-3. Molecular weight and L<strong>en</strong>nard-Jones parameters necessary to estimate diffusion coeffici<strong>en</strong>t [10]<br />

M i (g/mol) B k<br />

117<br />

ε (K) σ ) A<br />

o<br />

(<br />

CO 44 190 3.996<br />

2<br />

N 28 99.8 3.667<br />

2<br />

He 4 10.2 2.576<br />

Table 4-4. Estimation of diffusion coeffici<strong>en</strong>ts for binary gas mixtures He-CO2 and He-N2 at temperatures<br />

300, 350 and T = 323.<br />

7 °K, pressure 1 bar<br />

He − CO2<br />

2 N He −<br />

T (K) 300 350 323.7 300 350 323.7<br />

σ ( A)<br />

o<br />

12<br />

3.286 3.121<br />

ε / k (K) 44.02 31.90<br />

12<br />

*<br />

T (-) 6.815 7.950 7.353 9.403 10.970 10.145<br />

Ω D (-) 0.793 0.771 0.782 0.749 0.731 0.740<br />

D 12 ( cm s<br />

2<br />

) 0.596 0.772 0.677 0.715 0.925 0.812<br />

4.4.3 Effective diffusion coeffici<strong>en</strong>t in porous media<br />

A number of differ<strong>en</strong>t theoretical and experim<strong>en</strong>tal mo<strong>de</strong>ls have be<strong>en</strong> used to quantify gas<br />

diffusion processes in porous media. Most experim<strong>en</strong>tal mo<strong>de</strong>ls are mo<strong>de</strong>ls <strong>de</strong>rived for a<br />

free fluid (no porous media) that were modified for a porous medium. Attempts have be<strong>en</strong><br />

ma<strong>de</strong> to <strong>de</strong>fine effective diffusion parameters according to the pres<strong>en</strong>ce of the porous<br />

medium. In literature, the effective diffusion coeffici<strong>en</strong>ts are now well established,<br />

theoretically ([60], [104], [107], [90] and [79]) and experim<strong>en</strong>tally ([42], [22] and [49]).<br />

The comparison of the theoretical and experim<strong>en</strong>tal results for the <strong>de</strong>p<strong>en</strong>d<strong>en</strong>ce of the<br />

effective diffusion coeffici<strong>en</strong>t on the medium porosity shows that the results of Quintard<br />

(1993) in three dim<strong>en</strong>sional arrays of spheres [79] and the curve id<strong>en</strong>tified by Weissberg<br />

(1963) are in excell<strong>en</strong>t agreem<strong>en</strong>t with the experim<strong>en</strong>tal data [111].<br />

Many experim<strong>en</strong>tal studies have be<strong>en</strong> done to <strong>de</strong>termine the effective diffusion coeffici<strong>en</strong>t<br />

for unconsolidated porous media. The diffusion of hydrog<strong>en</strong> through cylindrical samples


of porous granular materials was measured by Currie (1960) [22]. An equation having two<br />

shape factor of the form<br />

*<br />

m<br />

D D = γ ε has be<strong>en</strong> proposed which fits with all granular<br />

material, m is the particle shape factor. The value of γ for glass spheres can be fixed to<br />

0.81 [15]. The expected m value for spheres is 1.5.<br />

For measuring effective diffusion coeffici<strong>en</strong>ts we have used the same apparatus and<br />

method but here, one part of the connecting tube (4 cm long, connected to the bottom bulb)<br />

is filled with the porous medium ma<strong>de</strong> of glass spheres (Fig. 4-8). A metal scre<strong>en</strong> was<br />

fixed at each <strong>en</strong>d of the tube to prev<strong>en</strong>t spheres fall down. The mesh size of the scre<strong>en</strong> is<br />

larger than spheres diameter and smaller than the pore size. The porosity of each medium<br />

has be<strong>en</strong> <strong>de</strong>termined by construction of a 3D image of the sample ma<strong>de</strong> with an X-ray<br />

tomography <strong>de</strong>vice (Skyscan 1174 type, see Fig. 4-9). A section image of the differ<strong>en</strong>t<br />

samples used in this study is shown in Fig. 4-10.<br />

A B C D<br />

Fig. 4-8. Cylindrical samples filled with glass sphere<br />

118


Glass spheres<br />

d= 700-1000 μm<br />

ε=42.5 %<br />

Mixture of glass spheres<br />

D=100-1000 μm<br />

ε=28.5 %<br />

Fig. 4-9. X-ray tomography <strong>de</strong>vice (Skyscan 1174 type) used in this study<br />

A B C D<br />

A B C<br />

Glass spheres<br />

d= 200-210 μm<br />

ε=40.2 %<br />

D E F<br />

Cylindrical material<br />

d=1500 μm<br />

ε=66 %<br />

119<br />

Glass spheres<br />

d= 100-125 μm<br />

ε=30.6 %<br />

Glass wool<br />

Mean fiber diameter= 6 μm<br />

ε ≅ 66 %<br />

Fig. 4-10. Section images of the tube (inner diameter d = 0.<br />

795cm)<br />

filled by differ<strong>en</strong>t materials obtained<br />

by an X-ray tomography <strong>de</strong>vice (Skyscan 1174 type)


The various diffusion time evolution through a free medium and porous media ma<strong>de</strong> of<br />

differ<strong>en</strong>t glass spheres (or mixture of them) are shown in Fig. 4-11 and Fig. 4-12 for He-<br />

N2 and He-CO2 systems, respectively. These results show clearly that the conc<strong>en</strong>tration<br />

time variations are very differ<strong>en</strong>t from free medium and porous medium experim<strong>en</strong>ts. In<br />

the case of porous medium there is a change <strong>de</strong>p<strong>en</strong>ding on the porosity of the medium.<br />

The values of the particle diameter, corresponding porosity, and calculated diffusion<br />

coeffici<strong>en</strong>ts are shown in Table 4-5 and Table 4-6. Here, the stared parameters are the<br />

effective coeffici<strong>en</strong>ts and the others are the coeffici<strong>en</strong>t in the free fluid.<br />

The diffusion coeffici<strong>en</strong>ts have be<strong>en</strong> obtained by curve fitting of equation ( 4-31) on the<br />

experim<strong>en</strong>tal data. We can conclu<strong>de</strong> from these results that there is not significant<br />

*<br />

differ<strong>en</strong>ce betwe<strong>en</strong> calculated ratios of D D12<br />

obtained from two differ<strong>en</strong>t gas systems.<br />

Conc<strong>en</strong>tration of N2 in bottom bulb (%) .<br />

100<br />

95<br />

90<br />

85<br />

80<br />

75<br />

70<br />

65<br />

60<br />

55<br />

50<br />

0 36000 72000 108000 144000 180000 216000<br />

Time (s)<br />

120<br />

Free Fluid<br />

Porous media, ε=42.55<br />

Porous media, ε=30.59<br />

Porous media, ε=28.52<br />

Fig. 4-11. Composition-time history in two-bulb diffusion cell for He-N2 system for differ<strong>en</strong>t medium.<br />

(<br />

0<br />

= 300K<br />

and c 100%<br />

)<br />

T C<br />

1 b =<br />

Table 4-5. Measured diffusion coeffici<strong>en</strong>t for He-N2 and differ<strong>en</strong>t media<br />

particle<br />

diameter (μm)<br />

Porosity<br />

(%)<br />

D12<br />

(cm 2 /s)<br />

D*/D12<br />

(-)<br />

Free Fluid 100 0.700 1<br />

750-1000 42.5 0.438 0.64<br />

100-125 30.6 0.397 0.57<br />

Mixture of spheres 28.5 0.355 0.51


Conc<strong>en</strong>tration of CO2 in bottom bulb (%)<br />

100<br />

95<br />

90<br />

85<br />

80<br />

75<br />

70<br />

65<br />

60<br />

55<br />

50<br />

0 72000 144000 216000 288000<br />

Time (s)<br />

121<br />

Free Fluid<br />

Porous media, ε=42.55<br />

Porous media, ε=40.21<br />

Porous media, ε=28.52<br />

Fig. 4-12. Composition-time history in two-bulb diffusion cell for He-CO2 system for differ<strong>en</strong>t medium<br />

0<br />

( T = 300K<br />

and c 100%<br />

)<br />

1 b =<br />

Table 4-6. Measured diffusion coeffici<strong>en</strong>t for He-CO2 and differ<strong>en</strong>t medium<br />

particle<br />

Porosity<br />

D<br />

diameter (μm) (%)<br />

(cm 2 D*/D<br />

/s)<br />

(-)<br />

Free Fluid 100 0.620 1<br />

750-1000 42.5 0.400 0.65<br />

200-210 40.2 0.375 0.61<br />

Mixture of spheres 28.5 0.322 0.52<br />

4.4.4 Free fluid and effective thermal diffusion coeffici<strong>en</strong>t<br />

Experim<strong>en</strong>tal investigations of thermal diffusion have usually be<strong>en</strong> based on the<br />

<strong>de</strong>termination of the differ<strong>en</strong>ce in composition of two parts of a giv<strong>en</strong> gas mixture which<br />

are at differ<strong>en</strong>t temperatures. In this work, after obtaining the steady-state in the diffusion<br />

process, <strong>de</strong>scribed in the section 4.4.3,(see also Fig. 4-13a), the temperature of top bulb is<br />

increased to 350°K as shown in Fig. 4-13b. In this stage the valve betwe<strong>en</strong> the two bulbs<br />

is closed. Increasing the temperature in this bulb will increase the pressure th<strong>en</strong>, by<br />

op<strong>en</strong>ing a tap on the top bulb, the pressure <strong>de</strong>crease until it reaches an equilibrium value<br />

betwe<strong>en</strong> the two bulbs.


a)<br />

b)<br />

T= 300K<br />

T=300K<br />

T= 350K<br />

T=300K<br />

t=0 t>0 t�∞<br />

Pure gas<br />

A<br />

Pure gas<br />

B<br />

Mixture<br />

A+B<br />

Mixture<br />

A+B<br />

Fig. 4-13. Schematic diagram of two bulb a) diffusion and b) thermal diffusion processes<br />

In this study, since the thermal diffusion coeffici<strong>en</strong>t is a complex function of<br />

conc<strong>en</strong>tration, temperature, pressure, and molecular masses of the compon<strong>en</strong>ts, we have<br />

tried to fix all these parameters in or<strong>de</strong>r to observe only the influ<strong>en</strong>ce of porosity on the<br />

thermal diffusion process.<br />

The separation can be found from the change in composition which occurs in one bulb<br />

during the experim<strong>en</strong>t, providing that the ratio of the volumes of the two bulbs is known.<br />

In the equation expressing the separation, the volume of the connecting tube has be<strong>en</strong><br />

neglected. Th<strong>en</strong>, from equation ( 4-18) the thermal diffusion factor for He-N2 and He-CO2<br />

binary mixtures is obtained, respectively, as about 0.31 and 0.36. In the literature, this<br />

factor, for the temperature range of 287°K -373 °K, is reported as about 0.36 for He-N2<br />

binary mixture [44]. From experim<strong>en</strong>tal results of composition <strong>de</strong>p<strong>en</strong>d<strong>en</strong>ce of He-CO2<br />

mixture, done with a swing separator method by Batabyal and Barua, (1968) [5] α T<br />

122<br />

Diffusion<br />

Thermal<br />

Diffusion<br />

Homog<strong>en</strong>ization<br />

Separation<br />

Mixture<br />

A+B<br />

Mixture<br />

A+B


increases with increasing conc<strong>en</strong>tration of the lighter compon<strong>en</strong>t. A thermal diffusion<br />

factor equal to 0.52 is obtained from the equation proposed in their paper, at T = 341.<br />

0 °K<br />

[5].<br />

A study using a two-bulb cell to <strong>de</strong>termine the composition and temperature <strong>de</strong>p<strong>en</strong>d<strong>en</strong>ce<br />

of the diffusion coeffici<strong>en</strong>t and thermal diffusion factor of He-CO2 system has be<strong>en</strong> done<br />

by Dunlop and Bignell, (1995) [28]. They obtain a diffusion coeffici<strong>en</strong>t equal to 0.615<br />

cm 2 /s at 300 °K and a thermal diffusion factor of 0.415 at T = 300 °K [28].<br />

The theoretical expression for the first approximation to the thermal diffusion factor,<br />

according to the Chapman-Enskog theory may be writt<strong>en</strong> as follows<br />

*<br />

[ ] = A(<br />

6C<br />

− 5)<br />

α ( 4-32)<br />

1<br />

12<br />

where A is a function of molecular weights, temperature, relative conc<strong>en</strong>tration of the two<br />

compon<strong>en</strong>ts, and *<br />

C 12 is a ratio of collision integrals in the principal temperature<br />

*<br />

<strong>de</strong>p<strong>en</strong>d<strong>en</strong>ce giv<strong>en</strong> by the ( 6C<br />

5)<br />

factor [40]. We calculated the values of the thermal<br />

12 −<br />

diffusion factors for He-CO2 and He-N2 mixtures at T , using this theoretical approach and<br />

according to the L<strong>en</strong>nard-Jones (12:6) pot<strong>en</strong>tial mo<strong>de</strong>l. We obtained a thermal diffusion<br />

factor for He-N2 mixture about 0.32 and for He-CO2 about 0.41. The <strong>de</strong>tail of formulation<br />

and estimation are listed in App<strong>en</strong>dix B.<br />

As we explained in section 4.2.4, wh<strong>en</strong> the experim<strong>en</strong>tal data concerning time evolution of<br />

the conc<strong>en</strong>tration exist, we can evaluate the both diffusion and thermal diffusion<br />

coeffici<strong>en</strong>ts. Therefore in this study, by a curve fitting procedure on the experim<strong>en</strong>tal data,<br />

two parameters 12 D and D T are adjusted until equation ( 4-26) fits the experim<strong>en</strong>tal curve.<br />

In fact, adjusting 12 D fits the slope of the experim<strong>en</strong>tal data curves and th<strong>en</strong>, D T is related<br />

to the final steady-state of the curves. The thermal diffusion kinetics for, respectively, free<br />

media and porous media ma<strong>de</strong> of differ<strong>en</strong>t glass spheres (or mixture of them) are shown in<br />

Fig. 4-14 (He-N2 mixture) and Fig. 4-15 (He-CO2 mixture). Here also, the time history<br />

changes with the porous medium porosity. The values for porosity, particle diameter of the<br />

porous medium, calculated diffusion coeffici<strong>en</strong>ts, thermal diffusion coeffici<strong>en</strong>ts and<br />

related thermal diffusion factor are shown in Table 4-7 for He-N2 mixture and Table 4-8<br />

for He-CO2 mixture.<br />

The diffusion coeffici<strong>en</strong>ts calculated with this method are larger than the one obtained in<br />

diffusion processes for He-N2 system. The theoretical approach (Table 4-4) and<br />

experim<strong>en</strong>tal data show that the diffusion coeffici<strong>en</strong>t increases with increasing the<br />

123


temperature. From equation (A. 2), wh<strong>en</strong> the i<strong>de</strong>al-gas law approximation is valid, we can<br />

write<br />

D<br />

12<br />

3<br />

2<br />

T<br />

∝ ( 4-33)<br />

*<br />

Ω D ( T )<br />

Conc<strong>en</strong>tration of N2 in bottom bulb (%)<br />

50.7<br />

50.6<br />

50.5<br />

50.4<br />

50.3<br />

50.2<br />

50.1<br />

50<br />

0 36000 72000 108000 144000 180000 216000 252000 288000<br />

124<br />

Time (s)<br />

Free Fluid<br />

Porous medium, ε=42.55<br />

Porous medium, ε=30.59<br />

Porous medium, ε=28.52<br />

Fig. 4-14. Composition-time history in two-bulb thermal diffusion cell for He-N2 binary mixture for<br />

0<br />

differ<strong>en</strong>t media ( Δ T = 50K<br />

, T = 323.<br />

7K<br />

and c 50%<br />

)<br />

1 = b<br />

Table 4-7. Measured thermal diffusion and diffusion coeffici<strong>en</strong>t for He-N2 and for differ<strong>en</strong>t media<br />

particle<br />

diameter<br />

(μm)<br />

Porosity<br />

(%)<br />

D12<br />

(cm 2 /s)<br />

D*/D12<br />

(-)<br />

αT<br />

(-)<br />

DT<br />

(cm 2 /s.K)<br />

Free Fluid 100 0.755 1 0.310 0.059 1<br />

DT*/DT<br />

(-)<br />

750-1000 42.5 0.457 0.605 0.312 0.035 0.61<br />

100-125 30.6 0.406 0.538 0.308 0.031 0.53<br />

Mixture of<br />

spheres<br />

28.5 0.369 0.489 0.304 0.028 0.48


Conc<strong>en</strong>tration of CO 2 in bottom bulb (%)<br />

50.7<br />

50.6<br />

50.5<br />

50.4<br />

50.3<br />

50.2<br />

50.1<br />

50<br />

0 36000 72000 108000 144000<br />

Time (s)<br />

125<br />

Free Fluid<br />

Porous medium, ε=42.55<br />

Porous medium, ε=40.21<br />

Porous medium, ε=28.52<br />

Fig. 4-15. Composition-time history in two-bulb thermal diffusion cell for He-CO2 binary mixture for<br />

0<br />

differ<strong>en</strong>t media .( Δ T = 50K<br />

, T = 323.<br />

7K<br />

and c 50%<br />

)<br />

1 b =<br />

Table 4-8. Measured diffusion coeffici<strong>en</strong>t and thermal diffusion coeffici<strong>en</strong>t for He-CO2 and for differ<strong>en</strong>t<br />

media<br />

particle<br />

diameter<br />

(μm)<br />

Porosity<br />

(%)<br />

D12<br />

(cm 2 /s)<br />

D*/D12<br />

(-)<br />

αT<br />

(-)<br />

DT<br />

(cm 2 /s.K)<br />

DT*/DT<br />

(-)<br />

Free Fluid 100 0.528<br />

1 0.358 0.047 1<br />

750-1000 42.5 0.304<br />

0.627 0.362 0.028 0.61<br />

200-210 40.2 0.320<br />

0.567 0.364 0.027 0.59<br />

Mixture of<br />

spheres<br />

28.5 0.273<br />

0.508 0.363 0.024 0.52<br />

In a second set of thermal diffusion experim<strong>en</strong>ts, we eliminate the valve betwe<strong>en</strong> the two<br />

bulbs in or<strong>de</strong>r to have a shorter relaxation time. In this case, the tube l<strong>en</strong>gth is equal to 4<br />

−4<br />

−2<br />

cm only (calibrated cell constant= 2.<br />

44×<br />

10 cm ) and we filled the system cells with a<br />

0<br />

binary gas mixture ( c 61.<br />

25%<br />

). At the initial state, the whole setup is kept at a<br />

N<br />

2 =<br />

uniform and constant temperature about 325 °K and the composition of the mixture is<br />

uniform everywhere. Th<strong>en</strong>, the temperature of the top bulb is increased to T = 350 °K<br />

H


and the temperature of the bottom bulb is lowered to T = 300 °K. At the <strong>en</strong>d of this<br />

process, wh<strong>en</strong> the temperature of each bulb remains constant, the pressure of the two bulbs<br />

is equal to the beginning of the experi<strong>en</strong>ce. The thermal diffusion separation in this period<br />

is very small because of the forced convection. The katharometer reading data have be<strong>en</strong><br />

recor<strong>de</strong>d with one minute interval. Conc<strong>en</strong>tration in bottom bulb has be<strong>en</strong> <strong>de</strong>termined<br />

using the katharometer calibration curve. Th<strong>en</strong>, with a curve fitting procedure on the<br />

experim<strong>en</strong>tal data, as in the last section, the two coeffici<strong>en</strong>ts 12 D and D T are adjusted until<br />

equation ( 4-26) fits the experim<strong>en</strong>tal curve. The adjusted curves for a free medium and<br />

differ<strong>en</strong>t porous media are shown in Fig. 4-17. The values obtained for porosity, particle<br />

diameter of the porous media, calculated diffusion and thermal diffusion coeffici<strong>en</strong>ts and<br />

thermal diffusion factor are listed in Table 4-9.<br />

Top<br />

bulb<br />

Bottom<br />

bulb<br />

Fig. 4-16. New experim<strong>en</strong>tal thermal diffusion setup without the valve betwe<strong>en</strong> the two bulbs<br />

126<br />

C<br />

Tube containing<br />

porous medium


Conc<strong>en</strong>tration of N2 in bottom bulb (%)<br />

61.8<br />

61.7<br />

61.6<br />

61.5<br />

61.4<br />

61.3<br />

61.2<br />

0 7200 14400 21600 28800 36000 43200 50400<br />

Time (s)<br />

127<br />

Free Fluid<br />

Porous medium (ε=33.78)<br />

Porous medium (ε=26.37)<br />

Fig. 4-17. Composition-time history in two-bulb thermal diffusion cell for He-N2 binary mixture for<br />

0<br />

differ<strong>en</strong>t media ( Δ T = 50K<br />

, T = 323.<br />

7K<br />

and c 61.<br />

25%<br />

)<br />

1 = b<br />

Table 4-9. Measured diffusion coeffici<strong>en</strong>t and thermal diffusion coeffici<strong>en</strong>t for He-N2 and differ<strong>en</strong>t media<br />

particle<br />

diameter<br />

(μm)<br />

Porosity<br />

(%)<br />

D<br />

(cm 2 /s)<br />

D*/D<br />

(-)<br />

αT<br />

(-)<br />

DT<br />

(cm 2 /s.K)<br />

Free Fluid - 0.480 1 0.256 0.029 1<br />

DT*/DT<br />

(-)<br />

315-325 33.8 0.257 0.530 0.248 0.015 0.52<br />

5-50 26.4 0.165 0.344 0.252 0.010 0.34<br />

4.4.5 Effect of solid thermal conductivity on thermal diffusion<br />

In section 2.7.2, the theoretical mo<strong>de</strong>l revealed that, for pure diffusion, the effective<br />

thermal diffusion coeffici<strong>en</strong>ts are in<strong>de</strong>p<strong>en</strong>d<strong>en</strong>t of the thermal conductivity ratio. To<br />

validate this result we have conducted some experim<strong>en</strong>ts with two differ<strong>en</strong>t materials<br />

shown in Fig. 4-18. Stainless steel and glass spheres are used in these experim<strong>en</strong>ts which<br />

their physical properties are listed in Table 4-10.


G H<br />

Fig. 4-18. Cylindrical samples filled with differ<strong>en</strong>t materials (H: stainless steal, G: glass spheres and<br />

ε=42.5)<br />

Table 4-10. The solid (spheres) and fluid mixture physical properties (T=300 K) [48]<br />

Material<br />

of<br />

particles<br />

Diameter<br />

(mm)<br />

k σ (sphere)<br />

(W/m.K)<br />

k β (gases)<br />

(W/m.K)<br />

( ρc p ) σ<br />

(kg/m3<br />

×J./kg.K)<br />

( ρc p ) β<br />

(kg/m3<br />

×J./kg.K)<br />

k σ<br />

kβ<br />

( ρcp<br />

)<br />

ρcp<br />

Stainless<br />

Steel<br />

1 15<br />

He=0.149<br />

7900×477<br />

He=<br />

0.1624×5200<br />

301 3202<br />

Glass<br />

1 1.1<br />

CO2=0.0181<br />

Mix=0.0499 2500×750<br />

CO2=<br />

1.788×844<br />

Mix=1177<br />

22 1593<br />

128<br />

σ<br />

( ) β<br />

The gas mixture used for this study is a He (50%)-CO2 (50%) mixture. Like in the section<br />

4.4.4, first the sample (tube of 4 cm l<strong>en</strong>gth filled by spheres produced a void fraction about<br />

ε=42.5) is placed betwe<strong>en</strong> two bulbs carefully. Next step is to vacuum the air from the<br />

system using the vacuum pump. Th<strong>en</strong>, the system, which is kept at a uniform temperature<br />

of 325 K, is filled by the gas mixture at atmospheric pressure. To start the thermal<br />

diffusion process, the temperature of the top bulb is increased to 350 K and at the same<br />

time, the temperature of the bottom bulb is <strong>de</strong>creased to 300 K. The advantage of this<br />

method is that, at the <strong>en</strong>d of this step, wh<strong>en</strong> the temperature of each bulb is constant, the<br />

pressure of the closed system will be the same as at the starting of the experim<strong>en</strong>t. During<br />

this intermediate period, the thermal diffusion process is negligible because of the forced


convection betwe<strong>en</strong> the two bulbs through the tube. Th<strong>en</strong>, continuous measurem<strong>en</strong>ts by<br />

katharometer, barometers, and thermometers are started in the two bulbs.<br />

It is important to note that, during the tube packing, ev<strong>en</strong> if the spheres diameter is the<br />

same in both cases, it may result in slightly differ<strong>en</strong>t porosity. This could be due to<br />

differ<strong>en</strong>t spheres arrangem<strong>en</strong>t (because of packing and shaking <strong>de</strong>gree). To avoid this<br />

error, we ma<strong>de</strong> three samples of each type therefore the experim<strong>en</strong>t is repeated for each<br />

sample.<br />

Katharometer differ<strong>en</strong>ce reading (mV)<br />

322<br />

320<br />

318<br />

316<br />

314<br />

312<br />

310<br />

0 36000 72000 108000 144000 180000<br />

Time (s)<br />

129<br />

Stainless steel<br />

Fig. 4-19. Katharometer reading time history in two-bulb thermal diffusion cell for He-CO2 binary mixture<br />

for porous media having differ<strong>en</strong>t thermal conductivity (3 samples of stainless steal and 3 samples of glass<br />

0<br />

spheres) ( Δ T = 50K<br />

, T = 323.<br />

7K<br />

and c 50%<br />

)<br />

1 = b<br />

Katharometer reading time histories for porous media ma<strong>de</strong> of differ<strong>en</strong>t thermal<br />

conductivity (stainless steal and glass spheres) is plotted in Fig. 4-19.<br />

As it is shown in this plot, thermal diffusion curves for the two differ<strong>en</strong>t materials can be<br />

superimposed and we can conclu<strong>de</strong> that, in this case, the thermal conductivity ratio has no<br />

significant influ<strong>en</strong>ce on the thermal diffusion process.<br />

Glas s


4.4.6 Effect of solid thermal connectivity on thermal diffusion<br />

It is known that heat conduction at the contact points plays a dominant role in <strong>de</strong>termining<br />

the effective heat conduction in porous media [79, 92]. To <strong>de</strong>termine the effect of solid<br />

phase connectivity, one must use a higher thermal conductor than stainless steal. Because<br />

theoretical results show that the influ<strong>en</strong>ce of this ph<strong>en</strong>om<strong>en</strong>on is consi<strong>de</strong>rable wh<strong>en</strong> the<br />

thermal conductivity ratio is more than 100 as shown in Fig. 2-15. Therefore, it is better to<br />

test more conductive martial. We have chos<strong>en</strong> aluminum and glass spheres with a diameter<br />

of 6mm shown in Fig. 4-20. The sample preparation is differ<strong>en</strong>t from the last section.<br />

Here, the sample is ma<strong>de</strong> of one array of spheres, in which we can neglect the problem that<br />

we had concerning spheres arrangem<strong>en</strong>t in the tube. Here, inner diameter of insulated<br />

rigid glass tube is chos<strong>en</strong> to be d = 0.<br />

75cm<br />

and l<strong>en</strong>gth of l = 5.<br />

8 cm. The number of<br />

spheres forming the porous medium is t<strong>en</strong>, which produced a void fraction about ε = 0.<br />

56 .<br />

Glass<br />

spheres<br />

Fig. 4-20. Cylindrical samples filled with differ<strong>en</strong>t materials (A: glass spheres, B: aluminium spheres and<br />

ε=0.56)<br />

The physical properties of two materials used in this experim<strong>en</strong>t are listed in Table 4-11.<br />

As we can see, the thermal conductivity ratio for aluminum and the mixture of helium and<br />

carbon dioxi<strong>de</strong> is about 4749. At this value, the connectivity of the solid phase has a high<br />

influ<strong>en</strong>ce on the effective thermal conductivity coeffici<strong>en</strong>ts as shown in Fig. 2-15.<br />

130<br />

Aluminium<br />

spheres<br />

10 mm


Table 4-11. The solid (spheres) and fluid mixture physical properties (T=300 K) [48]<br />

Material<br />

of<br />

particles<br />

Diameter<br />

(mm)<br />

k σ (sphere)<br />

(W/m.K)<br />

k β (gases)<br />

(W/m.K)<br />

( ρc p ) σ<br />

(kg/m3<br />

×J./kg.K)<br />

( ρc p ) β<br />

(kg/m3<br />

×J./kg.K)<br />

k σ<br />

kβ<br />

( ρcp<br />

)<br />

ρcp<br />

Aluminum 6 237 2702×903 4749 2073<br />

He=0.149<br />

Glass<br />

6 1.1<br />

CO2=0.0181<br />

Mix=0.0499 2500×750<br />

He=<br />

0.1624×5200<br />

CO2=<br />

1.788×844<br />

Mix=1177<br />

22 1593<br />

131<br />

σ<br />

( ) β<br />

Katharometer reading time histories for porous media ma<strong>de</strong> of aluminum and glass spheres<br />

have be<strong>en</strong> plotted in Fig. 4-21.<br />

This figure shows that, the thermal diffusion curves for two differ<strong>en</strong>t materials are<br />

superimposed. That means that, the particle-particle contact does not show a consi<strong>de</strong>rable<br />

influ<strong>en</strong>ce on the thermal diffusion process.<br />

Katharometer differ<strong>en</strong>ce reading (mV)<br />

322<br />

320<br />

318<br />

316<br />

314<br />

312<br />

310<br />

0 36000 72000 108000 144000 180000<br />

Time (s)<br />

Aluminium spheres<br />

Glass spheres<br />

Fig. 4-21. Katharometer time history in two-bulb thermal diffusion cell for He-CO2 binary mixture for<br />

porous media ma<strong>de</strong> of differ<strong>en</strong>t thermal conductivity (aluminum and glass spheres) ( Δ T = 50K<br />

,<br />

0<br />

T = 323.<br />

7K<br />

and c 50%<br />

)<br />

1 = b


4.4.7 Effect of tortuosity on diffusion and thermal diffusion coeffici<strong>en</strong>ts<br />

Mathematically, the tortuosity factor, τ , <strong>de</strong>fined as the ratio of the l<strong>en</strong>gth of the “tortuous”<br />

path in a porous media divi<strong>de</strong>d by a straight line value shown in Fig. 4-22.<br />

Fig. 4-22. Definition of tortuosity coeffici<strong>en</strong>t in porous media, L= straight line and L’= real path l<strong>en</strong>gth<br />

L'<br />

L<br />

There are several <strong>de</strong>finitions of this factor. The most wi<strong>de</strong>ly used correlation for gaseous<br />

diffusion is the one of Millington and Quirk (1961) for saturated unconsolidated system<br />

[63, 62]<br />

1 3<br />

τ = 1 ε<br />

( 4-34)<br />

Tortuosity is also an auxiliary quantity related to the ratio of the effective and free<br />

diffusion coeffici<strong>en</strong>ts. τ in many application for homog<strong>en</strong>ous and isotropic <strong>en</strong>vironm<strong>en</strong>t is<br />

also <strong>de</strong>fined as [69, 68]<br />

D12<br />

τ = *<br />

( 4-35)<br />

D<br />

In this study, we <strong>de</strong>fine the tortuosity as a ratio of the effective to free diffusion<br />

coeffici<strong>en</strong>ts as we m<strong>en</strong>tioned also in chapter 2<br />

*<br />

D 1 D<br />

= or,<br />

D τ D<br />

12<br />

*<br />

T<br />

T<br />

1<br />

⎛ L ⎞<br />

= ,which are f ⎜ ⎟ ( 4-36)<br />

τ<br />

⎝ L′<br />

⎠<br />

Table 4-12 pres<strong>en</strong>ts the tortuosity factors calculated from values measured in this work<br />

(for non-consolidate spheres and the tortuosity <strong>de</strong>finition with Eq. ( 4-36))<br />

132<br />

⎛ L ⎞<br />

τ<br />

= f ⎜ ⎟<br />

⎝ L′<br />


Table 4-12. Porous medium tortuosity coeffici<strong>en</strong>ts<br />

Particle<br />

diameter<br />

(μm)<br />

Porosity<br />

(%)<br />

τ =<br />

D<br />

D<br />

12<br />

*<br />

(From diffusion<br />

experim<strong>en</strong>ts)<br />

τ =<br />

133<br />

D<br />

D<br />

12<br />

*<br />

(From thermal diffusion<br />

experim<strong>en</strong>ts)<br />

τ =<br />

D<br />

D<br />

T<br />

*<br />

T<br />

(From thermal<br />

diffusion experim<strong>en</strong>ts)<br />

750-1000 42.5 1.57 1.62 1.63 1.61<br />

200-210 40.2<br />

315-325 33.8<br />

1.65 1.76 1.71 1.71<br />

- 1.89 1.92 1.90<br />

100-125 30.6 1.76 1.86 1.87 1.83<br />

Mixture of<br />

spheres<br />

28.5<br />

1.95 2.00 1.99 1.98<br />

5-50 26.4 - 2.91 2.90 2.90<br />

We showed that porosity has an important influ<strong>en</strong>ce on both effective isothermal diffusion<br />

and thermal diffusion coeffici<strong>en</strong>ts. Another question is what may happ<strong>en</strong> wh<strong>en</strong> there are<br />

two media with the same porosity but not the same tortuosity.<br />

In this section we tried to construct two media with such properties as shown in Fig. 4-23.<br />

E F<br />

Fig. 4-23. Cylindrical samples filled with differ<strong>en</strong>t materials producing differ<strong>en</strong>t tortuosity but the same<br />

porosity ε=66% (E: cylindrical material and F: glass wool)<br />

The section image of the tubes filled by these materials obtained by the tomograph <strong>de</strong>vice<br />

(Skyscan 1174 type) is shown in Fig. 4-10 E and F.<br />

The results of conc<strong>en</strong>tration-time histories for porous medium ma<strong>de</strong> with cylindrical<br />

samples and glass wool are plotted in Fig. 4-24. As we can see, the conc<strong>en</strong>tration-time<br />

curves for two cases are not superimposed and are completely separated. The calculated<br />

τ


tortuosity factor in porous medium ma<strong>de</strong> of cylindrical materials is τ = 2.<br />

37 and for glass<br />

wool it is relatively two times less than one for cylindrical materials τ = 1.<br />

04.<br />

These<br />

results indicate that the effective coeffici<strong>en</strong>ts are not only the function of porosity but also<br />

the geometry. Thus, tortuosity prediction using only the porosity may not be <strong>en</strong>ough and<br />

the permeability of the medium should be consi<strong>de</strong>red also.<br />

Conc<strong>en</strong>tration of N2 in bottom bulb (%)<br />

50.6<br />

50.5<br />

50.4<br />

50.3<br />

50.2<br />

50.1<br />

50<br />

0 18000 36000 54000 72000<br />

Time (s)<br />

134<br />

Cylindrical material<br />

Glass wool<br />

Free medium<br />

Fig. 4-24. Composition time history in two-bulb thermal diffusion cell for He-CO2 binary mixture in porous<br />

media ma<strong>de</strong> of the same porosity (ε=66% ) but differ<strong>en</strong>t tortuosity (cylindrical materials and glass wool)<br />

0<br />

( Δ T = 50K<br />

, T = 323.<br />

7K<br />

and c 50%<br />

)<br />

1 b =<br />

4.5 Discussion and comparison with theory<br />

In the theoretical part of this study, chapter 2, we have pres<strong>en</strong>ted the volume averaging<br />

method to obtain the macro-scale equations that <strong>de</strong>scribe diffusion and thermal diffusion<br />

processes in a homog<strong>en</strong>eous porous medium. The results of this mo<strong>de</strong>l showed that the<br />

effective thermal diffusion coeffici<strong>en</strong>t at diffusive regime can be estimated with the single<br />

tortuosity, results fully discussed in the literature [79, 80]. Here, we rewrite the basic<br />

theoretical results for a pure diffusion and binary system as<br />

D<br />

D<br />

*<br />

12<br />

*<br />

DT<br />

1<br />

= = , for pure diffusion ( 4-37)<br />

D τ<br />

T<br />

Fig. 4-25 and Fig. 4-26 show respectively a comparison of effective diffusion and<br />

thermal diffusion coeffici<strong>en</strong>ts measured in this study and the theoretical results from the<br />

volume averaging technique for differ<strong>en</strong>t porosity of the medium. We note that, the<br />

volume averaging process, have carried out using a mo<strong>de</strong>l unit cell such as the one shown


in Fig. 2-6. In this system, the effective diffusion and thermal diffusion coeffici<strong>en</strong>ts for<br />

differ<strong>en</strong>t fractional void space is plotted as the continuous lines. These figures show that<br />

the experim<strong>en</strong>tal, effective coeffici<strong>en</strong>t results for the non-consolidated porous media ma<strong>de</strong><br />

of spheres are in excell<strong>en</strong>t agreem<strong>en</strong>t with volume averaging theoretical estimation.<br />

D *<br />

ε<br />

D<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0.0<br />

Volume averaging, theoretical estimation<br />

Experim<strong>en</strong>tal, He-CO2<br />

Experim<strong>en</strong>tal, He-N2<br />

0.20 0.25 0.30 0.35 0.40 0.45 0.50 0.55 0.60<br />

ε (Porosity)<br />

Fig. 4-25. Comparison of experim<strong>en</strong>tal effective diffusion coeffici<strong>en</strong>t data with the theoretical one obtained<br />

from volume averaging technique for differ<strong>en</strong>t porosity and a specific unit cell<br />

In Fig. 4-27 the ratio of k T kT<br />

*<br />

thermal diffusion ratio, *<br />

k T , has be<strong>en</strong> <strong>de</strong>fined as<br />

has be<strong>en</strong> plotted against porosity, where, the effective<br />

135<br />

* * *<br />

kT = TDT<br />

D . The experim<strong>en</strong>tal results<br />

for both mixtures are fitted with the volume averaging theoretical estimation. These results<br />

also validate the theoretical results and reinforce the fact that for pure diffusion the Soret<br />

number is the same in the free medium and porous media.


D T<br />

*<br />

T<br />

ε<br />

D<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0.0<br />

Volume averaging, theoretical estimation<br />

Experim<strong>en</strong>tal, He-CO2<br />

Experim<strong>en</strong>tal, He-N2<br />

0.20 0.25 0.30 0.35 0.40 0.45 0.50 0.55 0.60<br />

ε (Porosity)<br />

Fig. 4-26. Comparison of experim<strong>en</strong>tal effective thermal diffusion coeffici<strong>en</strong>t data with theoretical one<br />

obtained from volume averaging technique for differ<strong>en</strong>t porosity and a specific unit cell<br />

k T<br />

*<br />

T<br />

k<br />

2.0<br />

1.8<br />

1.6<br />

1.4<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

Volume averaging, theoretical estimation<br />

Experim<strong>en</strong>tal, He-N2<br />

Experim<strong>en</strong>tal, He-CO2<br />

0.20 0.25 0.30 0.35 0.40 0.45 0.50 0.55 0.60<br />

ε (Porosity)<br />

Fig. 4-27. Comparison of the experim<strong>en</strong>tal thermal diffusion ratio data with theoretical one obtained from<br />

volume averaging technique for differ<strong>en</strong>t porosity and a specific unit cell<br />

136


4.6 Conclusion<br />

In this chapter, we used a “two-bulb apparatus” for measuring the diffusion and thermal<br />

diffusion coeffici<strong>en</strong>t in free medium and non-consolidated porous medium having differ<strong>en</strong>t<br />

porosity and thermal conductivity, separately. The results show that for He-N2 and He-CO2<br />

mixtures, the porosity of the medium has a great influ<strong>en</strong>ce on the thermal diffusion<br />

process. On the opposite, thermal conductivity and particle-particle contact of the solid<br />

phase have no significant influ<strong>en</strong>ce on thermal diffusion in porous media. The comparison<br />

of the ratio of effective coeffici<strong>en</strong>ts in the porous medium to the one in the free medium<br />

shows that the behavior of tortuosity is the same for the thermal diffusion coeffici<strong>en</strong>t and<br />

diffusion coeffici<strong>en</strong>t. Therefore, the thermal diffusion factor is the same for a free medium<br />

and porous media. For non-consolidated porous media ma<strong>de</strong> of the spheres these results<br />

agree with the mo<strong>de</strong>l obtained by upscaling technique for effective thermal diffusion<br />

coeffici<strong>en</strong>t proposed in the theoretical chapter 2. The tortuosity of the medium calculated<br />

using both effective diffusion and effective thermal diffusion coeffici<strong>en</strong>ts are not differ<strong>en</strong>t<br />

to the measurem<strong>en</strong>t accuracy.<br />

137


Chapter 5<br />

G<strong>en</strong>eral conclusions and perspectives


5. G<strong>en</strong>eral conclusions and perspectives<br />

In this study, the effective Darcy-scale coeffici<strong>en</strong>ts for coupled via Soret effect heat and<br />

mass transfer in porous media have be<strong>en</strong> <strong>de</strong>termined theoretically and experim<strong>en</strong>tally. A<br />

theoretical mo<strong>de</strong>l has be<strong>en</strong> <strong>de</strong>veloped using the volume averaging technique. We<br />

<strong>de</strong>termined from the microscopic equations new transport equations for averaged fields<br />

with some effective coeffici<strong>en</strong>ts. The associated quasi-steady closure problems related to<br />

the pore-scale physics have be<strong>en</strong> solved over periodic unit cells repres<strong>en</strong>tative of the<br />

porous structure. Particularly, we have studied the influ<strong>en</strong>ce of the void volume fraction<br />

(porosity), Péclet number and thermal conductivity on the effective thermal diffusion<br />

coeffici<strong>en</strong>ts. The obtained results show that<br />

• the values of the effective coeffici<strong>en</strong>ts in porous media are completely differ<strong>en</strong>t<br />

from the ones of the free medium (without the porous medium),<br />

• in all cases, the porosity of the medium has a great influ<strong>en</strong>ce on the effective<br />

thermal diffusion coeffici<strong>en</strong>ts,<br />

• for a diffusive regime, this influ<strong>en</strong>ce is the same for the effective diffusion<br />

coeffici<strong>en</strong>t and thermal diffusion coeffici<strong>en</strong>t. As a result, for low Péclet numbers,<br />

the effective Soret number in porous media is the same as the one in the free fluid.<br />

At this regime, the effective thermal diffusion coeffici<strong>en</strong>t does not <strong>de</strong>p<strong>en</strong>d on the<br />

solid to fluid conductivity ratio,<br />

• for a convective regime, the effective Soret number <strong>de</strong>creases and th<strong>en</strong> changes its<br />

sign. In this case, a change of conductivity ratio will change the effective thermal<br />

diffusion coeffici<strong>en</strong>t as well as the effective thermal conductivity coeffici<strong>en</strong>t,<br />

• theoretical results also showed that for pure diffusion, ev<strong>en</strong> if the effective thermal<br />

conductivity <strong>de</strong>p<strong>en</strong>ds on the particle-particle contact, the effective thermal<br />

diffusion coeffici<strong>en</strong>t is always constant and in<strong>de</strong>p<strong>en</strong>d<strong>en</strong>t on the connectivity of the<br />

solid phase.<br />

As a validation, the initial pore-scale problem was solved numerically over an array of<br />

cylin<strong>de</strong>rs, and the resulting averaged temperature and conc<strong>en</strong>tration fields were compared<br />

to macro-scale theoretical predictions using the effective coeffici<strong>en</strong>ts resulting from the<br />

previous theoretical study. The results showed that<br />

139


• a good agreem<strong>en</strong>t has be<strong>en</strong> found betwe<strong>en</strong> macro-scale resolutions and micro-<br />

scale, direct simulations, which validates the proposed theoretical mo<strong>de</strong>l,<br />

• thermal diffusion modifies the local conc<strong>en</strong>tration and this modification <strong>de</strong>p<strong>en</strong>ds<br />

locally on the porosity, thermal conductivity ratio and fluid velocity. Therefore, we<br />

cannot ignore this effect.<br />

A new experim<strong>en</strong>tal setup has be<strong>en</strong> <strong>de</strong>signed and ma<strong>de</strong>-up to <strong>de</strong>termine directly the<br />

effective diffusion and thermal diffusion coeffici<strong>en</strong>ts for binary mixtures. This setup is a<br />

closed system, which helped carry out the experim<strong>en</strong>ts for the case of pure diffusion only.<br />

The experim<strong>en</strong>ts have be<strong>en</strong> performed with special all-glass two-bulb apparatus,<br />

containing two double-spherical layers. The diffusion and thermal diffusion of heliumnitrog<strong>en</strong><br />

and helium-carbon dioxi<strong>de</strong> systems through cylindrical samples first without<br />

porous media and th<strong>en</strong> filled with spheres of differ<strong>en</strong>t diameters and thermal conductivities<br />

were measured at the atmospheric pressure. Conc<strong>en</strong>trations were <strong>de</strong>termined by analysing<br />

the gas mixture composition in the bulbs with a katharometer <strong>de</strong>vice. A transi<strong>en</strong>t-state<br />

method for coupled evaluation of thermal diffusion and Fick coeffici<strong>en</strong>t in two bulbs<br />

systems has be<strong>en</strong> proposed. Here, with a simple thermal diffusion experim<strong>en</strong>t, this mo<strong>de</strong>l<br />

is able to <strong>de</strong>termine both diffusion and thermal diffusion coeffici<strong>en</strong>ts. The <strong>de</strong>termination of<br />

diffusion and thermal diffusion coeffici<strong>en</strong>ts is done by a curve fitting of the temporal<br />

experim<strong>en</strong>tal results with the transi<strong>en</strong>t-state solution <strong>de</strong>scribing the mass balance betwe<strong>en</strong><br />

the two bulbs. The results showed<br />

• a <strong>de</strong>p<strong>en</strong>d<strong>en</strong>cy of the thermal diffusion and diffusion coeffici<strong>en</strong>ts on the porosity,<br />

• a good agreem<strong>en</strong>t with theoretical results, which confirm the validity of the<br />

theoretical results for pure diffusion,<br />

• the tortuosity of the medium calculated using both effective diffusion and effective<br />

thermal diffusion coeffici<strong>en</strong>ts were not differ<strong>en</strong>t to the measurem<strong>en</strong>t accuracy,<br />

• the experim<strong>en</strong>tal results also showed that the particle-particle touching has not a<br />

significant influ<strong>en</strong>ce on the effective thermal diffusion coeffici<strong>en</strong>ts.<br />

There is still much work to be done concerning thermal diffusion in porous media. Several<br />

perspectives can be proposed. The following ones pres<strong>en</strong>t especial interest<br />

• in the theoretical part of this study we <strong>de</strong>veloped a coupled heat and mass transfer<br />

macro-scale equation with a non-thermal equilibrium case (using a two-equation<br />

temperature problem). One may use this mo<strong>de</strong>l wh<strong>en</strong> the assumption of thermal<br />

equilibrium is not valid. However, the closure problems have not be<strong>en</strong> solved for<br />

140


this mo<strong>de</strong>l. Therefore, the next step may be to solve numerically these closure<br />

problems and th<strong>en</strong> compare with the one-equation results,<br />

• the mo<strong>de</strong>l proposed in this study is able to predict the effective coeffici<strong>en</strong>ts in a<br />

binary mixture of gas (or liquid) phase, the future works may be to focuse on more<br />

real and complex problems, i.e. multi-compon<strong>en</strong>t and multi-phase systems.<br />

• the effect of solid phase connectivity on the effective thermal conductivity and<br />

thermal diffusion coeffici<strong>en</strong>ts has be<strong>en</strong> investigated on a two dim<strong>en</strong>sional closure<br />

problem. In the case of thermal diffusion, we eliminated the particle touching in the<br />

y-direction to calculate the longitudinal thermal diffusion coeffici<strong>en</strong>t. In future<br />

work, one can resolve this problem using a three dim<strong>en</strong>sional mo<strong>de</strong>l, keeping the x,<br />

y and z touching parts as shown in Fig. 5-1. The three dim<strong>en</strong>sional mo<strong>de</strong>l may be<br />

also interesting for calculating effective thermal conductivity for purpose of<br />

comparison with earlier two dim<strong>en</strong>sional results,<br />

Fig. 5-1. 3D geometry of the closure problem with particle-particle touching ma<strong>de</strong> with COMSOL<br />

Multiphysics<br />

• by now, there is no qualitative agreem<strong>en</strong>t betwe<strong>en</strong> numerical and experim<strong>en</strong>tal<br />

results concerning separation in packed thermogravitational cell as shown in Fig.<br />

5-2 [31]. We showed that the ratio of thermal conductivity is very important for the<br />

convective regime. Therefore, this should change the separation rate in a packed<br />

thermogravitational cell. Therefore, it is very interesting to find the relation<br />

141


etwe<strong>en</strong> separation and thermal conductivity ratio. This may be achieved by simple<br />

micro-scale mo<strong>de</strong>lling or by the <strong>de</strong>sign of an experim<strong>en</strong>tal setup for a packed<br />

thermogravitational cell filled with differ<strong>en</strong>t materials. The results may reveal the<br />

reason of discrepancy which exists betwe<strong>en</strong> theoretical and experim<strong>en</strong>t results in a<br />

packed thermogravitational cell,<br />

Separation<br />

Rayleigh number<br />

Fig. 5-2. Discrepancy betwe<strong>en</strong> numerical results and experim<strong>en</strong>tal measurem<strong>en</strong>ts in a packed thermo-<br />

gravitational cell [31]<br />

• in our work the experim<strong>en</strong>ts have be<strong>en</strong> done for a non-consolidated material. The<br />

next experim<strong>en</strong>ts can be done using consolidated porous media. One will be able to<br />

produce two porous media of differ<strong>en</strong>t materials with exactly same porosity.<br />

• in this study using a katharometer <strong>de</strong>vice, the experim<strong>en</strong>ts have be<strong>en</strong> limited to<br />

binary systems. However, it is also important to measure directly the effective<br />

thermal diffusion coeffici<strong>en</strong>ts in ternary mixtures or beyond. In future work, using,<br />

for example, a gas chromatography <strong>de</strong>vice, the results will be ext<strong>en</strong><strong>de</strong>d to more<br />

than two compon<strong>en</strong>ts,<br />

• the experim<strong>en</strong>ts were performed for pure diffusion cases, which allowed us to<br />

validate the corresponding theoretical results only at pure diffusion. Designing<br />

another setup capable to measure the impact of dispersion on thermal diffusion (see<br />

Fig. 5-3) can be helpful to validate the theoretical results wh<strong>en</strong> the Péclet number<br />

is not zero, as well as for practical reasons.<br />

142


Temperature<br />

Conc<strong>en</strong>tration<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

Hot<br />

temperature<br />

0.2<br />

-0.2<br />

-0.4<br />

0<br />

0 2 4 6 8 10 12 14<br />

0<br />

0 2 4 6 8 10 12 14<br />

Fig. 5-3. Proposition of experim<strong>en</strong>tal setup for convective regime<br />

• Finally, it could be interesting to apply the new results obtained in this work to<br />

practical situations.<br />

143<br />

Cold<br />

temperature<br />

Thermometer<br />

Katharometer<br />

Flow Flow


Conclusions générales et perspectives <strong>en</strong> français<br />

Dans cette étu<strong>de</strong>, les coeffici<strong>en</strong>ts effectifs à l’échelle <strong>de</strong> Darcy pour le transfert couplé <strong>de</strong><br />

la chaleur et <strong>de</strong> la matière dans le milieu poreux ont été déterminés expérim<strong>en</strong>talem<strong>en</strong>t et<br />

théoriquem<strong>en</strong>t. Un modèle théorique a été développé <strong>en</strong> utilisant la métho<strong>de</strong> <strong>de</strong> prise <strong>de</strong><br />

moy<strong>en</strong>ne volumique. L'application du théorème <strong>de</strong> prise <strong>de</strong> moy<strong>en</strong>ne volumique sur les<br />

équations microscopiques décrivant les transports à l’échelle du pore permet d'obt<strong>en</strong>ir les<br />

nouvelles équations <strong>de</strong> transport pour les champs moy<strong>en</strong>s avec les coeffici<strong>en</strong>ts effectifs.<br />

<strong>Les</strong> problèmes <strong>de</strong> fermetures liées à la physique <strong>de</strong> l'échelle <strong>de</strong>s pores ont été résolus sur<br />

une cellule unitaire périodique représ<strong>en</strong>tative <strong>de</strong> la structure poreuse. En particulier, nous<br />

avons étudié l'influ<strong>en</strong>ce <strong>de</strong> la fraction volumique du pore (porosité), nombre <strong>de</strong> Péclet et<br />

<strong>de</strong> la conductivité thermique sur les coeffici<strong>en</strong>ts <strong>de</strong> thermodiffusion effectifs. <strong>Les</strong> résultats<br />

obt<strong>en</strong>us montr<strong>en</strong>t que :<br />

• les valeurs <strong>de</strong>s coeffici<strong>en</strong>ts effectifs <strong>en</strong> milieu poreux sont complètem<strong>en</strong>t différ<strong>en</strong>ts<br />

<strong>de</strong> celles du milieu libre (sans milieu poreux),<br />

• dans tous les cas, la porosité du milieu a une gran<strong>de</strong> influ<strong>en</strong>ce sur les coeffici<strong>en</strong>ts<br />

<strong>de</strong> thermodiffusion effectifs <strong>en</strong> milieu poreux,<br />

• pour un régime diffusif (Pe = 0), cette influ<strong>en</strong>ce est la même pour le coeffici<strong>en</strong>t <strong>de</strong><br />

diffusion isotherme effectif et le coeffici<strong>en</strong>t <strong>de</strong> thermodiffusion effectif. En<br />

conséqu<strong>en</strong>ce, pour les faibles nombres <strong>de</strong> Péclet, le nombre <strong>de</strong> Soret effectif dans<br />

le milieu poreux est le même que celui <strong>en</strong> milieu libre. Pour ce régime, le<br />

coeffici<strong>en</strong>t <strong>de</strong> thermodiffusion effectif ne dép<strong>en</strong>d pas du ratio <strong>de</strong>s conductivités<br />

thermiques,<br />

• pour le régime convectif (Pe ≠ 0), le nombre <strong>de</strong> Soret effectif diminue et change<br />

même <strong>de</strong> signe pour les régimes fortem<strong>en</strong>t convectifs. Dans ce cas, un changem<strong>en</strong>t<br />

du rapport <strong>de</strong> la conductivité thermique changera le coeffici<strong>en</strong>t thermodiffusion<br />

effectif ainsi que le coeffici<strong>en</strong>t conductivité thermique effective,<br />

• <strong>en</strong> diffusion pure, même si la conductivité thermique effective dép<strong>en</strong>d <strong>de</strong> la<br />

connectivité <strong>de</strong> phase soli<strong>de</strong>, le coeffici<strong>en</strong>t thermodiffusion effective est toujours<br />

constant et indép<strong>en</strong>dant <strong>de</strong> la connectivité <strong>de</strong> la phase soli<strong>de</strong>,<br />

Afin <strong>de</strong> vali<strong>de</strong>r les résultats théoriques précéd<strong>en</strong>ts, le problème d'échelle du pore a été<br />

résolu numériquem<strong>en</strong>t sur une série <strong>de</strong> cylindres. <strong>Les</strong> températures et conc<strong>en</strong>trations<br />

144


moy<strong>en</strong>nes sont comparées avec les prédictions macroscopiques <strong>en</strong> utilisant les coeffici<strong>en</strong>ts<br />

effectifs obt<strong>en</strong>us avec le modèle proposé. <strong>Les</strong> résultats montr<strong>en</strong>t que :<br />

• il y a un très bon accord <strong>en</strong>tre les résultats issus <strong>de</strong>s résolutions à l'échelle<br />

macroscopique (avec coeffici<strong>en</strong>ts effectifs) et microscopiques (simulations<br />

directes), ce qui vali<strong>de</strong> le modèle théorique proposé,<br />

• la thermodiffusion modifie la conc<strong>en</strong>tration locale et cette modification dép<strong>en</strong>d<br />

localem<strong>en</strong>t <strong>de</strong> la porosité, du ratio <strong>de</strong> conductivité thermique et <strong>de</strong> la vitesse du<br />

flui<strong>de</strong>. Par conséqu<strong>en</strong>t, cet effet ne peut pas être négligé dans la plupart <strong>de</strong>s cas.<br />

Un nouveau dispositif expérim<strong>en</strong>tal a été conçu et mis <strong>en</strong> place afin <strong>de</strong> déterminer<br />

directem<strong>en</strong>t les coeffici<strong>en</strong>ts <strong>de</strong> diffusion et thermodiffusion effectif pour <strong>de</strong>s mélanges<br />

binaires. Le dispositif réalisé est un système fermé, ce qui a permis d’effectuer <strong>de</strong>s<br />

expéri<strong>en</strong>ces pour les cas <strong>de</strong> diffusion pure. <strong>Les</strong> expéri<strong>en</strong>ces ont été réalisées avec un<br />

dispositif <strong>de</strong> type « <strong>de</strong>ux bulbes » spécifique, tout <strong>en</strong> verre, cont<strong>en</strong>ant une double couche<br />

sphérique permettant <strong>de</strong> contrôler les températures <strong>de</strong> chaque réservoir. La diffusion et la<br />

thermodiffusion <strong>de</strong> mélanges binaires hélium-azote, et d'hélium-dioxy<strong>de</strong> <strong>de</strong> carbone, à<br />

travers <strong>de</strong>s échantillons cylindriques d'abord sans milieux poreux, puis rempli avec <strong>de</strong>s<br />

sphères <strong>de</strong> différ<strong>en</strong>ts diamètres et <strong>de</strong> différ<strong>en</strong>tes conductivités thermiques est mesurée à<br />

pression atmosphérique. <strong>Les</strong> conc<strong>en</strong>trations sont déterminées <strong>en</strong> analysant la composition<br />

du mélange <strong>de</strong> gaz dans les ampoules à l’ai<strong>de</strong> d'un catharomètre qui est solidarisé à une<br />

partie <strong>de</strong> l'ampoule. Une métho<strong>de</strong> transitoire pour l'évaluation couplée du coeffici<strong>en</strong>t <strong>de</strong><br />

thermodiffusion et <strong>de</strong> diffusion <strong>de</strong> Fick dans le système <strong>de</strong> <strong>de</strong>ux ampoules a été proposée.<br />

Ici, avec une expéri<strong>en</strong>ce simple <strong>de</strong> thermodiffusion, ce modèle est capable <strong>de</strong> déterminer à<br />

la fois les coeffici<strong>en</strong>ts <strong>de</strong> diffusion et <strong>de</strong> thermodiffusion. La détermination <strong>de</strong> ces<br />

coeffici<strong>en</strong>ts est réalisé par ajustem<strong>en</strong>t (« fiting ») <strong>de</strong> la courbe expérim<strong>en</strong>tale <strong>de</strong> l’évolution<br />

temporelle <strong>de</strong>s conc<strong>en</strong>trations avec une solution analytique décrivant le bilan transitoire <strong>de</strong><br />

matière <strong>en</strong>tre les <strong>de</strong>ux ampoules. Cela permet d'ajuster les coeffici<strong>en</strong>ts jusqu'à ce que les<br />

équations se superpos<strong>en</strong>t avec les résultats expérim<strong>en</strong>taux. <strong>Les</strong> résultats ont montré :<br />

• une dép<strong>en</strong>dance <strong>de</strong>s coeffici<strong>en</strong>ts <strong>de</strong> thermodiffusion et <strong>de</strong> diffusion effectifs avec<br />

la porosité,<br />

• un bon accord avec les résultats théoriques, ce qui confirme la validité <strong>de</strong>s résultats<br />

théoriques <strong>en</strong> diffusion pure,<br />

145


• une valeur <strong>de</strong> la tortuosité du milieu id<strong>en</strong>tique lorsqu’elle est calculée à partir <strong>de</strong>s<br />

coeffici<strong>en</strong>ts <strong>de</strong> diffusion effectifs ou à l’ai<strong>de</strong> <strong>de</strong>s coeffici<strong>en</strong>ts <strong>de</strong> thermodiffusion<br />

effectifs, ce qui permet <strong>de</strong> proposer une valeur moy<strong>en</strong>ne du coeffici<strong>en</strong>t <strong>de</strong><br />

tortuosité du milieu,<br />

• le contact particule-particule n'a pas d’influ<strong>en</strong>ce significative sur les coeffici<strong>en</strong>ts <strong>de</strong><br />

thermodiffusion effectifs.<br />

Il reste <strong>en</strong>core beaucoup <strong>de</strong>s recherches et <strong>de</strong> développem<strong>en</strong>ts à faire concernant la<br />

thermodiffusion <strong>en</strong> milieux poreux. Plusieurs perspectives peuv<strong>en</strong>t être proposées; celles<br />

qui suiv<strong>en</strong>t prés<strong>en</strong>t<strong>en</strong>t un intérêt particulier <strong>en</strong> prolongem<strong>en</strong>t du travail réalisé :<br />

• dans la partie théorique <strong>de</strong> cette étu<strong>de</strong>, nous avons développé un modèle<br />

macroscopique décrivant <strong>de</strong> transfert <strong>de</strong> chaleur et matière avec une équation <strong>de</strong><br />

non-équilibre locale thermique (avec un problème à <strong>de</strong>ux équation pour<br />

température), qu’il peut être utiliser lorsque l'hypothèse <strong>de</strong> l'équilibre thermique<br />

n'est pas vali<strong>de</strong>. Toutefois, pour ce modèle les problèmes <strong>de</strong> fermeture n’ont pas<br />

été résolus. Une prochaine étape consisterait à résoudre numériquem<strong>en</strong>t ces<br />

problèmes <strong>de</strong>s fermetures et comparer <strong>en</strong>suite avec les résultats à une équation,<br />

• l'effet <strong>de</strong> la connectivité <strong>de</strong> la phase soli<strong>de</strong> sur les coeffici<strong>en</strong>ts <strong>de</strong> thermodiffusion<br />

et la conductivité thermique a été traité avec un problème <strong>de</strong> fermeture à <strong>de</strong>ux<br />

dim<strong>en</strong>sions. Dans le cas <strong>de</strong> la thermodiffusion, nous avons éliminé la connectivité<br />

<strong>de</strong> particules dans la direction y pour ne pas « bloquer » le transfert <strong>de</strong> matière et<br />

calculer le coeffici<strong>en</strong>t <strong>de</strong> thermodiffusion effective longitudinal. En utilisant un<br />

modèle à trois dim<strong>en</strong>sions, on pourrai résoudre ce problème <strong>en</strong> gardant la<br />

connectivité <strong>de</strong> la phase soli<strong>de</strong> dans les trois directions x, y et z (Fig. 5-1). Ce<br />

modèle <strong>en</strong> trois dim<strong>en</strong>sions pourrait d’ailleurs être égalem<strong>en</strong>t intéressant pour la<br />

détermination <strong>de</strong> la conductivité thermique effective,<br />

• jusqu’à prés<strong>en</strong>t, il n'y a pas d'accord <strong>en</strong>tre les résultats numériques et<br />

expérim<strong>en</strong>taux concernant la séparation dans un cellule thermogravitationelle [31]<br />

comme indiqué dans la Fig. 5 2. Nous avons montré que l’influ<strong>en</strong>ce du rapport <strong>de</strong>s<br />

conductivités thermiques est très importante pour le régime convectif. Ceci doit<br />

avoir <strong>de</strong>s conséqu<strong>en</strong>ces sur la séparation <strong>de</strong>s espèces obt<strong>en</strong>ue dans les cellules <strong>de</strong><br />

thermogravitation. Par conséqu<strong>en</strong>t, il serait intéressant <strong>de</strong> trouver cette influ<strong>en</strong>ce<br />

par modélisation numérique <strong>en</strong> l’échelle du pore ou par la réalisation d’un autre<br />

146


dispositif expérim<strong>en</strong>tal avec une cellule thermogravitationelle remplie <strong>de</strong> matériaux<br />

différ<strong>en</strong>ts. Ceci pourrait peut être montrer la raison <strong>de</strong>s diverg<strong>en</strong>ces qui exist<strong>en</strong>t<br />

<strong>en</strong>tre les résultats théoriques et expérim<strong>en</strong>taux dans la cellule<br />

thermogravitationelle,<br />

• dans ce travail les expéri<strong>en</strong>ces ont été réalisées pour <strong>de</strong>s matériaux non-consolidés.<br />

<strong>Les</strong> expéri<strong>en</strong>ces suivante peut-être faite <strong>en</strong> utilisant les milieux poreux consolidés.<br />

Certaines expéri<strong>en</strong>ces pourrai<strong>en</strong>t être réalisé pour déterminer par exemple l'impact<br />

<strong>de</strong> la conductivité thermique sur les coeffici<strong>en</strong>ts <strong>de</strong> thermodiffusion <strong>en</strong> milieu<br />

poreux avec différ<strong>en</strong>tes propriétés thermo-physiques,<br />

• dans cette étu<strong>de</strong>, l’utilisation d’un catharomètre, a limité les expéri<strong>en</strong>ces à <strong>de</strong>s<br />

systèmes binaires. Il serait égalem<strong>en</strong>t important <strong>de</strong> pouvoir mesurer directem<strong>en</strong>t les<br />

coeffici<strong>en</strong>ts <strong>de</strong> thermodiffusion effectifs dans <strong>de</strong>s mélanges ternaires. Dans les<br />

travaux futurs, l’utilisation par exemple d’un dispositif <strong>de</strong> chromatographie seront<br />

permettrait d’obt<strong>en</strong>ir <strong>de</strong>s résultats pour <strong>de</strong>s mélanges à plus <strong>de</strong> <strong>de</strong>ux composants,<br />

• le dispositif expérim<strong>en</strong>tal réalisé ici a permis <strong>de</strong> traiter le cas <strong>de</strong> la diffusion pure,<br />

et <strong>de</strong> vali<strong>de</strong>r les résultats théoriques correspondant à ce cas. La validation<br />

expérim<strong>en</strong>tale <strong>de</strong>s résultats théoriques lorsque le nombre <strong>de</strong> Péclet n'est pas nul<br />

(validés par ailleurs par les simulations à l’échelle du pore) nécessiterait la<br />

réalisation d’un nouveau dispositif (voir Fig. 5-3), <strong>en</strong> système ouvert, permettant<br />

<strong>de</strong> mesurer l'impact <strong>de</strong> la dispersion sur la thermodiffusion.<br />

• Enfin, il pourrait être intéressant d'appliquer les nouveaux résultats obt<strong>en</strong>us dans ce<br />

travail à <strong>de</strong>s situations pratiques.<br />

147


App<strong>en</strong>dix A. Estimation of the diffusion coeffici<strong>en</strong>t with gas kinetic theory<br />

The diffusion coeffici<strong>en</strong>t D 12 for the isothermal diffusion of species 1 through constant-<br />

pressure binary mixture of species 1 and 2 is <strong>de</strong>fined by the relation<br />

J = −D<br />

∇c<br />

(A. 1)<br />

1<br />

12<br />

1<br />

where 1 J is the flux of species 1 and 1<br />

c is the conc<strong>en</strong>tration of the diffusing species.<br />

Mutual-diffusion, <strong>de</strong>fined by the coeffici<strong>en</strong>t D 12 , can be viewed as diffusion of species 1 at<br />

infinite dilution through species 2, or equival<strong>en</strong>tly, diffusion of species 2 at infinite<br />

dilution through species 2.<br />

Self-diffusion, <strong>de</strong>fined by the coeffici<strong>en</strong>t D 11,<br />

is the diffusion of a substance through itself.<br />

There are differ<strong>en</strong>t theoretical mo<strong>de</strong>ls for computing the mutual and self diffusion<br />

coeffici<strong>en</strong>t of gases. For non-polar molecules, L<strong>en</strong>nard-Jones pot<strong>en</strong>tials provi<strong>de</strong> a basis for<br />

computing diffusion coeffici<strong>en</strong>ts of binary gas mixtures [76]. The mutual diffusion<br />

coeffici<strong>en</strong>t, in units of cm 2 /s is <strong>de</strong>fined as<br />

D<br />

12<br />

3 2 M1<br />

+ M 2 1<br />

= 0.<br />

00188T<br />

2<br />

M M pσ<br />

Ω<br />

1<br />

2<br />

where T is the gas temperature in unit of Kelvin, 1 M and 2<br />

12<br />

D<br />

148<br />

M are molecular weights of<br />

(A. 2)<br />

species 1 and 2, p is the total pressure of the binary mixture in unit of bar, σ 12 is the<br />

L<strong>en</strong>nard-Jones characteristic l<strong>en</strong>gth, <strong>de</strong>fined by σ 1 2(<br />

σ + σ )<br />

12 = 1 2 , D<br />

Ω is the collision<br />

integral for diffusion, is a function of temperature, it <strong>de</strong>p<strong>en</strong>ds upon the choice of the<br />

intermolecular force law betwe<strong>en</strong> colliding molecules. Ω D is tabulated as a function of<br />

*<br />

the dim<strong>en</strong>sionless temperature T = kBT<br />

ε12<br />

for the 12-6 L<strong>en</strong>nard-Jones pot<strong>en</strong>tial, k B is the<br />

Boltzman gas constant and ε 12 = ε1ε<br />

2 is the maximum attractive <strong>en</strong>ergy betwe<strong>en</strong> two<br />

molecules. The accurate relation of Neufield et al. (1972) is<br />

1.<br />

06036 0.<br />

19300 1.<br />

03587 1.<br />

76474<br />

Ω D = +<br />

+<br />

+<br />

(A. 3)<br />

* 0.<br />

15610<br />

*<br />

*<br />

*<br />

( T ) exp(<br />

0.<br />

47635T<br />

) exp(<br />

1.<br />

52996T<br />

) exp(<br />

3.<br />

89411T<br />

)<br />

Values of the parameters σ and ε are known for many substances [76].<br />

The self-diffusion coeffici<strong>en</strong>t of a gas can be obtained from Eq. (A. 2), by observing that<br />

for a one-gas system: M M = M<br />

D<br />

12<br />

3 2 2 1<br />

= 0.<br />

00188T<br />

2<br />

M pσ<br />

Ω<br />

1 = 2 , ε 1 = ε 2 and 1 σ 2<br />

11<br />

D<br />

σ = . Thus,<br />

(A. 4)


App<strong>en</strong>dix B. Estimation of the thermal diffusion factor with gas kinetic theory<br />

From the kinetic theory of gases, the thermal diffusion factor, α T for a binary gas mixture<br />

is very complex, as <strong>de</strong>scribed by Chapman and Cowling, 1939. Three differ<strong>en</strong>t theoretical<br />

expressions for α T are available, <strong>de</strong>p<strong>en</strong>ding on the approximation procedures employed:<br />

the first approximation and second one of Chapman and Cowling and the first<br />

approximation of Kihara, 1949. The most accurate of these is probably Chapman and<br />

Cowling’s second approximation, but this is rather complicated, and the numerical<br />

computation involved is quite annoying. A few sample calculations indicated that Kihara’s<br />

expression is more accurate than Chapman and Cowling’s first approximation (Mason and<br />

Rice, 1954; Mason, 1954), and usually differs from their second approximation by less<br />

than the scatter in differ<strong>en</strong>t experim<strong>en</strong>tal <strong>de</strong>termination of α T . It therefore seemed<br />

satisfactory for the pres<strong>en</strong>t purpose to use Kihara’s approximation writt<strong>en</strong> in the form<br />

[ ] ( 6 5)<br />

*<br />

⎛ S1x1<br />

− S ⎞<br />

2x<br />

2<br />

α = ⎜<br />

⎟ C −<br />

T<br />

1<br />

⎜<br />

⎝ Q x<br />

2<br />

1 1<br />

+ Q x<br />

2<br />

2<br />

2<br />

+ Q<br />

12<br />

x x<br />

1<br />

2<br />

⎟<br />

⎠<br />

12<br />

149<br />

(B. 1)<br />

The principal contribution to the temperature <strong>de</strong>p<strong>en</strong>d<strong>en</strong>ce of α T comes from the factor<br />

( 6 5)<br />

*<br />

C , which involves only the unlike (1, 2) molecular interaction. The conc<strong>en</strong>tration<br />

12 −<br />

<strong>de</strong>p<strong>en</strong>d<strong>en</strong>ce is giv<strong>en</strong> by S1x1 − S2x<br />

2 term. The main <strong>de</strong>p<strong>en</strong>d<strong>en</strong>ce on the masses of the<br />

molecules is giv<strong>en</strong> by 1 S and 2 S . A positive value of α T means that compon<strong>en</strong>t 1 t<strong>en</strong>ds to<br />

move into the cooler region and 2 towards the warmer region. The temperature at which<br />

the thermal diffusion factor un<strong>de</strong>rgoes a change of sign is referred to as the inversion<br />

temperature.<br />

These quantities calculated as<br />

S<br />

1<br />

Q<br />

1<br />

M<br />

=<br />

M<br />

=<br />

M<br />

1<br />

2<br />

2<br />

2M<br />

2<br />

M + M<br />

2<br />

1<br />

2<br />

( M + M )<br />

1<br />

⎡⎛<br />

5 6<br />

× ⎢⎜<br />

− B<br />

⎣⎝<br />

2 5<br />

2<br />

*<br />

12<br />

⎞<br />

⎟M<br />

⎠<br />

⎡Ω<br />

⎢<br />

⎣ Ω<br />

( 2,<br />

2)<br />

11<br />

11<br />

2<br />

*<br />

2 12<br />

2<br />

2<br />

( 1,<br />

1)<br />

*<br />

σ ( M + M )<br />

12<br />

*<br />

2M<br />

2<br />

M + M<br />

2<br />

1<br />

1<br />

+ 3M<br />

⎤⎛<br />

σ<br />

⎥ ⎜<br />

⎦⎝<br />

2<br />

2<br />

2<br />

12<br />

+<br />

⎞<br />

⎟<br />

⎠<br />

⎡Ω<br />

⎢<br />

⎣ Ω<br />

( 2,<br />

2)<br />

11<br />

( 1,<br />

1)<br />

12<br />

8<br />

5<br />

4M<br />

1M<br />

−<br />

M<br />

*<br />

*<br />

1<br />

1<br />

⎤⎛<br />

σ<br />

⎥ ⎜<br />

⎦⎝<br />

σ<br />

M<br />

2<br />

11<br />

12<br />

A<br />

*<br />

12<br />

⎞<br />

⎟<br />

⎠<br />

2<br />

⎤<br />

⎥<br />

⎦<br />

A<br />

15 M<br />

−<br />

2<br />

( )<br />

( ) 2<br />

2 M 2 − M 1<br />

M + M<br />

1<br />

2<br />

(B. 2)<br />

(B. 3)


Q<br />

12<br />

⎛ M<br />

= 15 ⎜<br />

⎝ M<br />

1<br />

1<br />

−<br />

+<br />

M<br />

M<br />

⎛ 12<br />

× ⎜11−<br />

B<br />

⎝ 5<br />

2<br />

2<br />

*<br />

12<br />

2<br />

⎞ ⎛ 5 6<br />

⎟ ⎜ − B<br />

⎠ ⎝ 2 5<br />

⎞ 8<br />

⎟ +<br />

⎠ 5<br />

( M + M )<br />

( 2,<br />

2)<br />

( M + M ) ⎡Ω<br />

1<br />

M<br />

2<br />

*<br />

12<br />

M<br />

⎞ 4M<br />

1M<br />

⎟ +<br />

⎠<br />

1<br />

2<br />

⎢<br />

⎣ Ω<br />

1<br />

11<br />

( 1,<br />

1)<br />

12<br />

*<br />

*<br />

A<br />

*<br />

2 12<br />

2<br />

2<br />

⎤⎡Ω<br />

⎥⎢<br />

⎦⎣<br />

Ω<br />

150<br />

( 2,<br />

2)<br />

22<br />

( 1,<br />

1)<br />

12<br />

*<br />

*<br />

⎤⎛<br />

σ σ<br />

⎥⎜<br />

⎜<br />

⎦⎝<br />

σ<br />

11 22<br />

2<br />

12<br />

⎞<br />

⎟<br />

⎠<br />

2<br />

(B. 4)<br />

with relations for S2, Q2 <strong>de</strong>rived from S1, Q1 by interchange of subscripts. The transport<br />

properties for gaseous mixtures can also be expressed in terms of the same collision<br />

integral.<br />

A<br />

B<br />

C<br />

Ω<br />

*<br />

A 12 ,<br />

( 2,<br />

2)*<br />

*<br />

12 = ( 1,<br />

1)*<br />

Ω<br />

*<br />

12<br />

5Ω<br />

=<br />

Ω<br />

( 1,<br />

2)*<br />

( 1,<br />

2)*<br />

*<br />

12 = ( 1,<br />

1)*<br />

Ω<br />

*<br />

B 12 and<br />

− 4Ω<br />

( 1,<br />

1)*<br />

Ω<br />

The subscripts on the<br />

( 1,<br />

3)*<br />

*<br />

*<br />

C 12 are function of 12 kT ε12<br />

T = <strong>de</strong>fined as<br />

(B. 5)<br />

(B. 6)<br />

(B. 7)<br />

( l,<br />

s)*<br />

Ω refer to the three differ<strong>en</strong>t binary molecular interactions which<br />

may occur in a binary gas mixture. By conv<strong>en</strong>tion, the subscript 1 refers to the heavier<br />

gas. To this investigation, L<strong>en</strong>nard-Jones (12-6) mo<strong>de</strong>l is applied, which has be<strong>en</strong> the best<br />

intermolecular pot<strong>en</strong>tial used to date for the study of transport ph<strong>en</strong>om<strong>en</strong>a and is expressed<br />

by a repulsion term varying as the inverse twelfth power of the distance of separation<br />

betwe<strong>en</strong> the c<strong>en</strong>ters of two molecules and an attraction term varying as the sixth power of<br />

the separation distance. The force constants of pure compon<strong>en</strong>ts σ and ε obtained from<br />

viscosity data are used as Table 4-3 and Table 4-4.


Refer<strong>en</strong>ces<br />

[1] A. Ahmadi, A. Aigueperse, and M. Quintard. Calculation of the effective properties<br />

<strong>de</strong>scribing active dispersion in porous media: from simple to complex unit cells. Advances<br />

in Water Resources, 24 (3-4):423–438, 2000.<br />

[2] A. Ahmadi, M. Quintard, and S. Whitaker. Transport in chemically and mechanically<br />

heterog<strong>en</strong>eous porous media V. Two-equation mo<strong>de</strong>l for solute transport with adsorption.<br />

Advances in water resources, 22:59–58, 1998.<br />

[3] A. Amiri and K. Vafai. Analysis of dispersion effects and non-thermal equilibrium,<br />

non-darcian, variable porosity incompressible flow through porous media. International<br />

journal of heat and mass transfer, 37(6):939–954, 1994.<br />

[4] P. S. Arora, I. R. Shankland, T. N. Bell, M. A. Yabsley, and P. J. Dunlop. Use of<br />

precise binary diffusion coeffici<strong>en</strong>ts to calibrate two-bulb cells instead of using the<br />

standard <strong>en</strong>d correction for the connecting tube. Review of Sci<strong>en</strong>tific Instrum<strong>en</strong>t, 48<br />

(6):673–674, 1977.<br />

[5] A. K. Batabyal and A. K. Barua. Composition <strong>de</strong>p<strong>en</strong>d<strong>en</strong>ce of the thermal-diffusion<br />

factors in He-CO2 Ne-CO2, and Xe-CO2 mixtures. Journal of Chemical Physics, 48:2557–<br />

2560, 1968.<br />

[6] J.C. Batsale, C. Gobbé, and M. Quintard. Local non-equilibrium heat transfer in porous<br />

media, in rec<strong>en</strong>t research <strong>de</strong>velopm<strong>en</strong>ts in heat, mass & mom<strong>en</strong>tum transfer. Research<br />

Signpost, India 1:1–24, 1996.<br />

[7] J. Bear. Hydraulics of Groundwater. McGraw-Hill Book Company, New York, 1979.<br />

[8] E. W. Becker. The thermal <strong>de</strong>-mixing of gases at high pressures. Die<br />

Naturwiss<strong>en</strong>schaft<strong>en</strong>, 37 (7):165–166, 1950.<br />

[9] A. Bejan and D. A. Nield. Convection in Porous Media. Springer Verlag, 1998.<br />

[10] R.B. Bird, W.E. Stewart, and E.N. Lightfoot. Transport Ph<strong>en</strong>om<strong>en</strong>a. 2nd Ed. New<br />

York: J. Wiley and Sons, 2002.<br />

[11] H. Br<strong>en</strong>ner. Dispersion resulting from flow through spatially periodic porous media.<br />

Philosophical Transactions of the Royal Society of London. Series A, Mathematical and<br />

Physical Sci<strong>en</strong>ces, 297 (1430):81–133, 1980.<br />

[12] H. Brown. On the temperature assignm<strong>en</strong>ts of experim<strong>en</strong>tal thermal diffusion<br />

coeffici<strong>en</strong>ts. Physical Review, 58:661 – 662, 1940.<br />

151


[13] D. R. Caldwell. Thermal and Fickian diffusion of sodium chlori<strong>de</strong> in a solution of<br />

oceanic conc<strong>en</strong>tration. Deep-Sea Research and Oceanographic Abstracts, 20 (11):1029–<br />

1039, 1973.<br />

[14] R. G. Carbonell and S. Whitaker. Heat and mass transfer in porous media.<br />

Fundam<strong>en</strong>tals of Transport Ph<strong>en</strong>om<strong>en</strong>a in Porous Media, Martinus Nijhoff, Dordrecht,<br />

1984.<br />

[15] P. C. Carman. Flow of gases through porous media. Aca<strong>de</strong>mic Press Inc., New York,<br />

1956.<br />

[16] M. Chan<strong>de</strong>sris and D. Jamet. Jump conditions and surface-excess quantities at a<br />

fluid/porous interface: A multi-scale approach. Transport in Porous Media, 78:419–438,<br />

2009.<br />

[17] S. Chapman and T. G. Cowling. Thermal Combustion and Diffusion in Gases, An<br />

Account of the Kinetic Theory of Viscosity, Thermal Conduction and Diffusion in Gases.<br />

Cambridge Mathematical Library, third Ed., 1970.<br />

[18] M. C. Charrier-Mojtabi, K. Maliwan, Y. Pedramrazi, G. Bardan, and A. Mojtabi.<br />

Control of thermoconvective flows by vibration. Mecanique and Industries, 4 (5):545–<br />

549, 2003.<br />

[19] F. Cherblanc, A. Ahmadi, and M. Quintard. Two-domain <strong>de</strong>scription of solute<br />

transport in heterog<strong>en</strong>eous porous media: Comparison betwe<strong>en</strong> theoretical predictions and<br />

numerical experim<strong>en</strong>ts. Advances in Water Resources, 30 (5):1127–1143, 2007.<br />

[20] P. Costeseque, T. Pollak, J. K. Platt<strong>en</strong>, and M. Marcoux. Transi<strong>en</strong>t-state method for<br />

coupled evaluation of Soret and Fick coeffici<strong>en</strong>ts, and related tortuosity factors, using free<br />

and porous packed thermodiffusion cells. European Physical Journal E, Soft matter, 15<br />

(3):249–253, 2004.<br />

[21] P. Costeseque, D. Fargue, and Ph. Jamet. Thermodiffusion in porous media and its<br />

consequ<strong>en</strong>ces. Thermal Nonequilibrium Ph<strong>en</strong>om<strong>en</strong>a in Fluid Mixtures. Lecture Notes in<br />

Physics, Springer, Berlin, 584:389–427, 2002.<br />

[22] J. A. Currie. Gaseous diffusion in porous media. I. A non-steady state method. British<br />

Journal of Applied Physics, 11 (8):314–317, 1960.<br />

[23] J. H. Cushman, L. S. B<strong>en</strong>nethum, and B. X. Hu. A primer on upscaling tools for<br />

porous media. Advances In Water Resources, 25(8-12):1043–1067, AUG-DEC 2002.<br />

[24] H. Davarzani, J. Chastanet, M. Marcoux, and M. Quintard. Theoretical <strong>de</strong>termination<br />

of effective thermodiffusion coeffici<strong>en</strong>ts, application to the <strong>de</strong>scription of mass transfer in<br />

152


porous media. Lecture Notes of the 8th International Meeting on Thermodiffusion (IMT8),<br />

Thermal Nonequilibrium,, 3, Forschungsz<strong>en</strong>trum jülich GmbH, Bonn, Germany:181–187,<br />

2008.<br />

[25] H. Davarzani, M. Marcoux, and M. Quintard. Theoretical predictions of the effective<br />

thermodiffusion coeffici<strong>en</strong>ts in porous media. International journal of heat and mass<br />

transfer, 53: 1514-1528, 2010.<br />

[26] H. A. Daynes. Gas analysis by measurem<strong>en</strong>t of thermal conductivity. Cambridge<br />

University Press, Cambridge, UK, pages 1–302, 1933.<br />

[27] H. A. Daynes and G. A. Shakespear. The theory of the katharometer. Proceedings of<br />

the Royal Society of London. Series A, Containing Papers of a Mathematical and Physical<br />

Character, 97, Issue 685:273–286, 1920.<br />

[28] P. J. Dunlop and C. M. Bignell. Diffusion coeffici<strong>en</strong>ts and thermal diffusion factors<br />

for He-CO2, He-N2O and He-COS. Berichte <strong>de</strong>r Buns<strong>en</strong>-Gesellschaft, 99:77–79, 1995.<br />

[29] R. E. Ewing. Aspects of upscaling in simulation of flow in porous media. Advances in<br />

Water Resources, 20 (5-6):349–358, 1997.<br />

[30] D. Fargue, P. Costeseque, P. Ph. Jamet, and S. Girard-Gaillard. Separation in vertical<br />

temperature gradi<strong>en</strong>t packed thermodiffusion cells: an unexpected physical explanation to<br />

a controversial experim<strong>en</strong>tal problem. Chemical Engineering Sci<strong>en</strong>ce, 59 (24):5847–5852,<br />

2004.<br />

[31] D. Fargue, P. Costeseque, P. Ph. Jamet, and S. Girard-Gaillard. Separation in vertical<br />

temperature gradi<strong>en</strong>t packed thermodiffusion cells: an unexpected physical explanation to<br />

a controversial experim<strong>en</strong>tal problem. Chemical Engineering Sci<strong>en</strong>ce, 59 (24):5847–5852,<br />

2004.<br />

[32] A. Firoozabadi. Thermodynamics of Hydrocarbon Reservoirs. McGraw-Hill, New<br />

York City, 1991.<br />

[33] E.N. Fuller, K. Ensley, and J.C. Giddings. Diffusion of halog<strong>en</strong>ated hydrocarbons in<br />

helium: the effect of structure on collision cross sections. Journal of Physical Chemistry,<br />

73:3679, 1969.<br />

[34] G. Galliéro, J. Colombani, P. A. Bopp, B. Duguay, J. P. Caltagirone, and F. Montel.<br />

Thermal diffusion in micropores by molecular dynamics computer simulations. Physica A:<br />

Statistical Mechanics and its Applications, 361 (2):494–510, 2006.<br />

153


[35] Ph. Georis, F. Montel, S. Van Vaer<strong>en</strong>bergh, Y. Decroly, and J. C. Legros.<br />

Measurem<strong>en</strong>t of the Soret coeffici<strong>en</strong>t in cru<strong>de</strong> oil. In Proceedings of the European<br />

Petroleum Confer<strong>en</strong>ce, 1998.<br />

[36] M. Giglio and A. V<strong>en</strong>dramini. Thermal l<strong>en</strong>s effect in a binary liquid mixture: a new<br />

effect. Applied Physics Letters, 25 (10):555–557, 1974.<br />

[37] K.E. Grew and T.L. Ibbs. Thermal diffusion in gases. Cambridge University Press,<br />

1952.<br />

[38] S.R. De Groot. Non-Equilibrium Thermodynamics. Dover Publication Inc., New<br />

York, 1984.<br />

[39] S.R. De Groot and P. Mazur. Non-Equilibrium Thermodynamics. Dover, New York,<br />

1984.<br />

[40] J. O. Hirschfel<strong>de</strong>r, C.F. Curtiss, and R. B. Bird. Molecular Theory of Gases and<br />

Liquids. Johns Wiley & Sons, Inc., New York, 1964.<br />

[41] Clifford K. Ho and S.W. Webb (Eds). Gas transport in porous media. Springer<br />

(Online service), 2006.<br />

[42] J. Hoogschag<strong>en</strong>. Diffusion in porous catalysts and adsorb<strong>en</strong>ts. Industrial and<br />

Engineering Chemistry, 47 (5):906–912, 1955.<br />

[43] W. Hort, S. J. Linz, and M. Lücke. Onset of convection in binary gas mixtures: role of<br />

the dufour effect. Physical review. A, 45(6):3737–3748, 1992.<br />

[44] T. L. Ibbs and K. E. Grew. Influ<strong>en</strong>ce of low temperatures on the thermal diffusion<br />

effect. Proceedings of the Physical Society, 43:142–156, 1931.<br />

[45] I. Ryzhkov Ilya. On double diffusive convection with Soret effect in a vertical layer<br />

betwe<strong>en</strong> co-axial cylin<strong>de</strong>rs. Physica D 215, 215 (2):191–200, 2006.<br />

[46] G. Jessop. Katharometers. Journal of Sci<strong>en</strong>tific Instrum<strong>en</strong>ts, 43 (11):777–782, 1966.<br />

[47] M. Kaviany. Principles of Heat Transfer in Porous Media. Second Edition, (Second<br />

Printing), Springer-Verlag, New York, 1999.<br />

[48] M. Kaviany. Principles of Heat Transfer. John Wiley & Sons, Inc., New York, 2002.<br />

[49] J. H. Kim, J. A. Ochoa, and S. Whitaker. Diffusion In Anisotropic Porous-Media.<br />

Transport In Porous Media, 2:327–356, 1987.<br />

[50] L. S. Kotousov and A. V. Panyushkin. Apparatus for <strong>de</strong>termining thermal-diffusion<br />

coeffici<strong>en</strong>ts in gas mixtures. Journal of Engineering Physics and Thermophysics, 9<br />

(5):390–393, 1965.<br />

154


[51] B. Lacabanne, S. Blancher, R. Creff, and F. Montel. Soret effect in multicompon<strong>en</strong>t<br />

flow through porous media: local study and upscaling process. Lecture notes in physics,<br />

Springer, 584:448–465, 2002.<br />

[52] L. D. Landau and E. M. Lifschitz. Fluid Mechanics. 2nd Ed. Oxford, England:<br />

Pergamon Press, 1982.<br />

[53] D. Lasseux, A. Ahmadi, and A. A. Abbasian Arani. Two-phase inertial flow in<br />

homog<strong>en</strong>eous porous media: A theoretical <strong>de</strong>rivation of a macroscopic mo<strong>de</strong>l. Transport<br />

in Porous Media, 75 (3):371–400, 2008.<br />

[54] M. Lor<strong>en</strong>z and A.H. Emery. The packed thermodiffusion column. Chemical<br />

Engineering Sci<strong>en</strong>ce, 11:16–23, 1959.<br />

[55] C. Ludwig. Sitzungsber. K. Preuss, Akad. Wiss., 20:539, 1856.<br />

[56] M. Marcoux and M. C. Charrier-Mojtabi. Etu<strong>de</strong> paramétrique <strong>de</strong> la thermogravitation<br />

<strong>en</strong> milieu poreux. Comptes R<strong>en</strong>dus <strong>de</strong> l’Académie <strong>de</strong>s Sci<strong>en</strong>ces, 326:539–546, 1998.<br />

[57] M. Marcoux and P. Costeseque. Study of transversal dim<strong>en</strong>sion influ<strong>en</strong>ce on species<br />

separation in thermogravitational diffusion columns. Journal of Non-Equilibrium<br />

Thermodynamics, 32 (3):289–298, 2007.<br />

[58] E. A. Mason and S. C. Sax<strong>en</strong>a. Approximate formula for the thermal conductivity of<br />

gas mixtures. Physics of Fluids, 1:361–369, 1958.<br />

[59] E.A. Mason and A.P. Malinauskas. Gas Transport in Porous Media: The Dusty Gas<br />

Mo<strong>de</strong>l. Elsevier, Amsterdam, 1983.<br />

[60] J. C. Maxwell. A treatise on electricity and magnetism. Clar<strong>en</strong>don Press, Oxford, I,<br />

2nd edition, 1881.<br />

[61] C. C. Mei. Method of homog<strong>en</strong>ization applied to dispersion in porous media.<br />

Transport in porous media, 9(3):261–274, 1991.<br />

[62] R. J. Millington. Gas diffusion in porous media. Sci<strong>en</strong>ce, 130 (3367):100–102, 1959.<br />

[63] R. J. Millington. Permeability of porous solids. Transactions of the Faraday Society,<br />

57:1200–1207, 1961.<br />

[64] C. Moyne. Two-equation mo<strong>de</strong>l for a diffusive process in porous media using the<br />

volume averaging method with an unsteady-state closure. Advances in Water Resources,<br />

20, Issues 2-3:63–76, 1997.<br />

[65] C. Moyne, S. Didierjean, H.P.A. Souto, and O.T. da Silveira Filho. Thermal<br />

dispersion in porous media: One-equation mo<strong>de</strong>l. International Journal of Heat and Mass<br />

Transfer, 43 (20):3853–3867, 2000.<br />

155


[66] H. Nasrabadi, H. Hoteit, and A. Firoozabadi. An analysis of species separation in<br />

thermogravitational column filled with porous media. Transport in Porous Media, 67<br />

(3):473–486, 2007.<br />

[67] E. P. Ney and F. C. Armistead. The self-diffusion coeffici<strong>en</strong>t of uranium<br />

hexafluori<strong>de</strong>. Physical Review, 71(1):14–19, 1947.<br />

[68] C. Nicholson. Diffusion and related transport mechanisms in brain tissue. Reports on<br />

Progress in Physics, 64:815–884, 2001.<br />

[69] C. Nicholson and J. M. Phillips. Ion diffusion modified by tortuosity and volume<br />

fraction in the extracellular micro<strong>en</strong>vironm<strong>en</strong>t of the rat cerebellum. Journal of Physiology<br />

(Cambridge), 321:225–258, 1981.<br />

[70] D. A. Nield and A. Bejan. Convection in Porous Media. Second Edition, Springer-<br />

Verlag New York, Inc., New York, 1999.<br />

[71] J. Ochoa-Tapia and S. Whitaker. Heat transfer at the boundary betwe<strong>en</strong> a porous<br />

medium and a homog<strong>en</strong>eous fluid. International Journal of Heat and Mass Transfer, 40<br />

(11):2691–2707, 1997.<br />

[72] J. K. Platt<strong>en</strong>. Enhanced molecular separation in inclined thermogravitational columns.<br />

The Journal of Physical Chemistry B, 107 (42):11763–11767, 2003.<br />

[73] J. K. Platt<strong>en</strong>. The Soret effect: A review of rec<strong>en</strong>t experim<strong>en</strong>tal results. Journal of<br />

applied mechanics, 73(1):5–15, 2006.<br />

[74] J. K. Platt<strong>en</strong> and P. Costeseque. The Soret coeffici<strong>en</strong>t in porous media. Journal of<br />

Porous Media, 7 (4):329–42, 2004.<br />

[75] J.K. Platt<strong>en</strong> and J.C. Legros. Convection in liquids. Springer, Berlin, Chap. 9, 1984.<br />

[76] B. E. Poling, J. M. Rausnitz, and J. P. Connell. The Properties of Gases and Liquids.<br />

McGraw-Hill, New York, 5th edition, 2000.<br />

[77] N. Puiroux, M. Prat, and M. Quintard. Non-equilibrium theories for macroscale heat<br />

transfer: ablative composite layer systems. International Journal of Thermal Sci<strong>en</strong>ces,<br />

43(6):541–554, 2004.<br />

[78] M. Quintard, and S. Whitaker. Transport in or<strong>de</strong>red and disor<strong>de</strong>red porous media III:<br />

Closure and comparison betwe<strong>en</strong> theory and experim<strong>en</strong>t. Transport in Porous Media,<br />

15(1):31–49, 1994.<br />

[79] M. Quintard. Diffusion in isotropic and anisotropic porous systems: Threedim<strong>en</strong>sional<br />

calculations. Transport in Porous Media, Volume 11, Number 2:187–199,<br />

1993.<br />

156


[80] M. Quintard, L. Bletzaker, D. Ch<strong>en</strong>u, and S. Whitaker. Nonlinear, multicompon<strong>en</strong>t,<br />

mass transport in porous media. Chemical Engineering Sci<strong>en</strong>ce, 61:2643–2696, 2006.<br />

[81] M. Quintard, F. Cherblanc, and S. Whitaker. Dispersion in heterog<strong>en</strong>eous porous<br />

media: One-equation non-equilibrium mo<strong>de</strong>l. Transport in Porous Media, 44(1):181–203,<br />

2001.<br />

[82] M. Quintard, M. Kaviany, and S. Whitaker. Two-medium treatm<strong>en</strong>t of heat transfer in<br />

porous media: numerical results for effective properties. Advances in Water Resources, 20<br />

(2-3):77–94, 1997.<br />

[83] M. Quintard and S. Whitaker. Transport in or<strong>de</strong>red and disor<strong>de</strong>red porous media:<br />

volume-averaged equations, closure problems and comparison with experim<strong>en</strong>t. Chemical<br />

Engineering Sci<strong>en</strong>ce, 14:2534–2537, 1993.<br />

[84] M. Quintard and S. Whitaker. Local thermal equilibrium for transi<strong>en</strong>t heat<br />

conduction: theory and comparison with numerical experim<strong>en</strong>ts. International Journal of<br />

Heat and Mass Transfer, 38 (15):2779–2796, 1995.<br />

[85] M. Quintard and S. Whitaker. Theoretical analysis of transport in porous media.<br />

Handbook of Heat Transfer in Porous Media, edited by H. Hadim and K. Vafai, Marcel<br />

Decker, Inc., New York,, Ch. 1:1–52, 2000.<br />

[86] G. D. Rabinovich, R. Y. Gurevich, and G. N. Bobrova. Thermodiffusion separation of<br />

liquid mixtures. [in Russian], Nauka i Tekhnika, Minsk, 1971.<br />

[87] G.D. Rabinovich. Separation of isotopes and other mixtures by thermal diffusion.<br />

Atomizdat, Moscow, 1981.<br />

[88] J. W. S. Rayleigh. Theory of Sound, (2nd Ed.). Dover, New-York, 1945.<br />

[89] D. Reith and F. Mueller-Plathe. On the nature of thermal diffusion in binary l<strong>en</strong>nardjones<br />

liquids. Journal of Chemical Physics, 112:2436–2443, 2000.<br />

[90] D. Ryan, R. G. Carbonell, and S. A. Whitaker. Theory of diffusion and reaction in<br />

porous media. A.I.Ch.E. Symposium Series, 77, No. 202:46–62, 1981.<br />

[91] M.Z. Saghir, C.G. Jiang, M. Chacha, Y. Yan, M. Khawaja, and S. Pan.<br />

Thermodiffuison in porous media. In Transport Ph<strong>en</strong>om<strong>en</strong>a in Porous Media III, (Eds. D.<br />

B. Ingham and I. Pop,, Elsevier, Oxford, page 227–260, 2005.<br />

[92] M. Sahraoui and M. Kaviany. Slip and no–slip temperature boundary conditions at<br />

interface of porous, plain media: Conduction. International Journal of Heat and Mass<br />

Transfer, 36:1019–1033, 1993.<br />

157


[93] M. Sahraoui and M. Kaviany. Direct simulation vs volume-averaged treatm<strong>en</strong>t of<br />

adiabatic, premixed flame in a porous medium. International Journal of Heat and Mass<br />

Transfer, 37(18):2817–2834, 1994.<br />

[94] E. Sanchez-Pal<strong>en</strong>cia. On the asymptotics of the fluid flow past an array of fixed<br />

obstacles. International Journal of Engineering Sci<strong>en</strong>ce, 20 (12):1291–1301, 1982.<br />

[95] S. C. Sax<strong>en</strong>a and E. A. Mason. Thermal diffusion and the approach to the steady state<br />

in gases II. Molecular Physics, 2 (4):379–396, 1959.<br />

[96] M. E. Schimpf. Studies in thermodiffusion by thermal field-flow fractionation.<br />

Proceedings of the Third International Symposium on Thermodiffusion, pages 58–63,<br />

1998.<br />

[97] A. G. Shashkov, A. F. Zolotukhina, T N. Abram<strong>en</strong>ko, and B. P. Mathur. Thermal<br />

diffusion factors for binary gas systems: Ar-N2, Ar-CO2, He-H2, He-N2O, Kr-N2O and He-<br />

NH3. Journal of physics B- Atomic molecular and optical physics, 12(21):3619–3630,<br />

1979.<br />

[98] C. Soret. Influ<strong>en</strong>ce <strong>de</strong> la temperature sur la distribution <strong>de</strong>s sels dans leurs solutions.<br />

Compte-R<strong>en</strong>du <strong>de</strong> l’Aca<strong>de</strong>mie <strong>de</strong>s Sci<strong>en</strong>ces, Paris, 91:289, 1880.<br />

[99] C. Tropea and and J. F. Foss (Eds.) A. L. Yarin. Springer Handbook of Experim<strong>en</strong>tal<br />

Fluid Mechanics. Springer, 2007.<br />

[100] S. Van Vaer<strong>en</strong>bergh, J. C. Legros, J. L. Daridon, T. Karapantsios, M. Kostoglou, and<br />

Z. M. Saghir. Multicompon<strong>en</strong>t transport studies of cru<strong>de</strong> oils and asphalt<strong>en</strong>es in DSC<br />

program. Microgravity Sci<strong>en</strong>ce and Technology, 18 (3-4):150–154, 2006.<br />

[101] A. van Itterbeek and O. van Paemel. Measurem<strong>en</strong>ts on the viscosity of hydrog<strong>en</strong>and<br />

<strong>de</strong>uterium gas betwe<strong>en</strong> 293°k and 14°k. Physica, 5 (10):938–944, 1938.<br />

[102] A. van Itterbeek and O. van Paemel. Measurem<strong>en</strong>ts on the viscosity of neon,<br />

hydrog<strong>en</strong>, <strong>de</strong>uterium and helium as a function of the temperature, betwe<strong>en</strong> room<br />

temperature and liquid hydrog<strong>en</strong> temperatures. Physica., 7 (3):265–272, 1940.<br />

[103] A. S. M. Wahby and J. Los. Diffusion in lor<strong>en</strong>tzian and quasi-lor<strong>en</strong>tzian N2-light<br />

noble gas mixtures. Physica, B + C, 145 (1):69–77, 1987.<br />

[104] N. Wakao and J. M. Smith. Diffusion in catalyst pellets. Chemical Engineering<br />

Sci<strong>en</strong>ce, 17 (11):825–834, 1962.<br />

[105] L. Wang and M. Quintard. Nanofluids of the future. Advances in Transport<br />

Ph<strong>en</strong>om<strong>en</strong>a, chap. 4:179–243, 2009.<br />

158


[106] H. Watts. Diffusion of krypton-85 in multicompon<strong>en</strong>t mixtures of krypton with<br />

helium, neon, argon and x<strong>en</strong>on. Transactions of the Faraday Society, 60 (10):1745–1751,<br />

1964.<br />

[107] H. L. Weissberg. Effective diffusion coeffici<strong>en</strong>t in porous media. Journal of Applied<br />

Physics, 34 (9):2636–2639, 1963.<br />

[108] K. R. Weller, N. S. St<strong>en</strong>house, and H. Watts. Diffusion of gases in porous solids. II.<br />

theoretical background and experim<strong>en</strong>tal method. Canadian Journal of Chemistry, 52<br />

(15):2684–2691, 1974.<br />

[109] S. Whitaker. Diffusion and dispersion in porous media. AIChE J, 13:420–427, 1967.<br />

[110] S. Whitaker. The forchheimer equation: A theoretical <strong>de</strong>velopm<strong>en</strong>t. Transport in<br />

Porous Media, 25 (1):27–61, 1996.<br />

[111] S. Whitaker. The method of Volume Averaging. Kluwer Aca<strong>de</strong>mic Publishers,<br />

Dordrecht, The Netherlands, 1999.<br />

[112] S. Wiegand. Thermal diffusion in liquid mixtures and polymer solutions. Journal of<br />

Physics: Cond<strong>en</strong>sed Matter, 16 (10):357–379, 2004.<br />

[113] S. Wiegand and W. Köhler. Measureemt of transport coeffici<strong>en</strong>ts by an optical<br />

grating technique. Thermal Nonequilibrium Ph<strong>en</strong>om<strong>en</strong>a in Fluid Mixtures Lecture Notes<br />

in Physics, Springer-Verlag, 584:189–210, 2002.<br />

[114] M. A. Yabsley and P. J. Dunlop. A study of the two-bulb method for measuring<br />

diffusion coeffici<strong>en</strong>ts of binary gas mixtures. Physica A, 85 (1):160–174, 1976.<br />

[115] F. Zanotti and R. G. Carbonell. Developm<strong>en</strong>t of transport equations for multiphase<br />

system-I. G<strong>en</strong>eral <strong>de</strong>velopm<strong>en</strong>t for two phase system. Chemical Engineering Sci<strong>en</strong>ce, 39,<br />

Issue 2:263–278, 1984.<br />

[116] F. Zanotti and R. G. Carbonell. Developm<strong>en</strong>t of transport equations for multiphase<br />

systems-II. Application to one-dim<strong>en</strong>sional axi-symmetric flows of two phases. Chemical<br />

Engineering Sci<strong>en</strong>ce, 39, Issue 2:279–297, 1984.<br />

159

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

Saved successfully!

Ooh no, something went wrong!