31.07.2015 Views

Diffuse interface models in fluid mechanics

Diffuse interface models in fluid mechanics

Diffuse interface models in fluid mechanics

SHOW MORE
SHOW LESS

Create successful ePaper yourself

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

is no longer the case and some of them are now discussed.At a critical po<strong>in</strong>t, the properties of both phases become equal. Above the critical po<strong>in</strong>t, atwo-phase system cannot exist and the <strong>fluid</strong> exists only as a one-phase <strong>fluid</strong>. On the contrary, belowthe critical po<strong>in</strong>t, under certa<strong>in</strong> conditions, the <strong>fluid</strong> can co-exist under two different phasesseparated by an <strong><strong>in</strong>terface</strong>. As the two-phase system system approaches the critical po<strong>in</strong>t frombelow, the thickness of the <strong><strong>in</strong>terface</strong> actually <strong>in</strong>creases and becomes <strong>in</strong>f<strong>in</strong>ite at the critical po<strong>in</strong>t.Thus, just below the critical po<strong>in</strong>t, the typical size of the transition layer that separates the bulkphases becomes of the same order of magnitude as the typical size of the bulk phases. Therefore,the <strong><strong>in</strong>terface</strong> cannot be modeled as a surface of discont<strong>in</strong>uity and its <strong>in</strong>ternal structure has to bedescribed .Let us consider two air bubbles <strong>in</strong> liquid water. It is commonly observed that two bubbles cancoalesce, i.e. they merge to give rise to a s<strong>in</strong>gle bubble. If the <strong><strong>in</strong>terface</strong>s are modeled as a surfaceof discont<strong>in</strong>uity, just at the moment where they merge, the model is s<strong>in</strong>gular. In a sense, one ofthe <strong><strong>in</strong>terface</strong>s desappears and this is not possible if the <strong><strong>in</strong>terface</strong> is discont<strong>in</strong>uous. To overcomethis s<strong>in</strong>gularity, one has to study the detailed <strong>in</strong>teraction of the <strong><strong>in</strong>terface</strong>s dur<strong>in</strong>g their merg<strong>in</strong>g.This can be done only by account<strong>in</strong>g for the <strong>in</strong>ternal structure of the <strong><strong>in</strong>terface</strong>s [Lee et al., 2002a,b].Another situation, actually similar to the previous one, is the description of the creation of asecond phase <strong>in</strong> an <strong>in</strong>itially s<strong>in</strong>gle-phase system; this is called nucleation. In this case, one hasto describe how, from a s<strong>in</strong>gle phase, an <strong><strong>in</strong>terface</strong> is created. This cont<strong>in</strong>uous process can bemodeled only if the <strong><strong>in</strong>terface</strong> under construction is modeled as a cont<strong>in</strong>uous medium [Dell’Isolaet al., 1996].1.3.2 Numerical limitationsIn the previous section, we have shown that, <strong>in</strong> some cases, the idealization of an <strong><strong>in</strong>terface</strong> asa surface of discont<strong>in</strong>uity is physically irrelevant and the detailed structure has to be described.However, <strong>in</strong> many applications, these phenomena can be neglected or do not occur. Nevertheless,the coupled partial differential equations that describe the two-phase flow are highlynon-l<strong>in</strong>ear and numerical simulation is often necessary to solve them. The numerical simulationof two-phase flows is very challeng<strong>in</strong>g because it is a mov<strong>in</strong>g boundary problem. Several numericaltechniques exist to solve this k<strong>in</strong>d of problems and their description is beyond the scopeof this presentation. These techniques are often difficult to implement numerically, especially <strong>in</strong>three space dimensions, and sometimes depend on the know-how of the code developper. Thisis partially due to the lack of a clear mathematical background for some of the methods. Nevertheless,the boundary conditions that must be applied at the mov<strong>in</strong>g <strong><strong>in</strong>terface</strong>s need a particulartreatment <strong>in</strong> the numerical algorithm, which is difficult and often tedious. This is why diffuse<strong><strong>in</strong>terface</strong> methods can be numerically attractive. If one can come up with a system of partial differentialequations that is valid <strong>in</strong> the entire two-phase system, <strong>in</strong>clud<strong>in</strong>g with<strong>in</strong> the cont<strong>in</strong>uoustransition <strong>in</strong>terfacial zones, the motion of the entire two-phase system would be describe by thiss<strong>in</strong>gle system of equations, which thus elim<strong>in</strong>ates the difficult problem of the particular treatmentof the boundary conditions at the <strong><strong>in</strong>terface</strong>s. The programm<strong>in</strong>g effort would therefore be highlydecreased. Moreover, if these equations are obta<strong>in</strong>ed from first pr<strong>in</strong>ciples, the development ofaccurate numerical schemes can be based on a better mathematical ground [Jamet et al., 2002].2 Liquid-vapor flows with phase-change: the van der Waals modelof capillarityIn this section, we present the van der Waals model of capillarity. This model is a diffuse <strong><strong>in</strong>terface</strong>model dedicated to the description of an <strong><strong>in</strong>terface</strong> that separates a liquid and a vapor phase of8


a pure <strong>fluid</strong>. Extensions to b<strong>in</strong>ary mixtures is possible but will not be presented here [Fouilletet al., 2002]. It is <strong>in</strong>terest<strong>in</strong>g to note that this model is the first diffuse <strong><strong>in</strong>terface</strong> developed byvan der Waals [van der Waals, 1894].2.1 Thermodynamic modelAny diffuse <strong><strong>in</strong>terface</strong> model is actually a thermodynamic model. Indeed, the <strong>in</strong>ternal structureof an <strong><strong>in</strong>terface</strong> is ma<strong>in</strong>ly an equilibrium feature. Dynamic effects only perturb this equilibriumstructure, which is thus important to characterize.2.1.1 A mean-field approximationThe ma<strong>in</strong> issue is the follow<strong>in</strong>g: is it possible to describe the <strong>in</strong>ternal structure of a liquid-vapor<strong><strong>in</strong>terface</strong> at equilibrium by consider<strong>in</strong>g a “classical” thermodynamic description of the <strong>fluid</strong>? By“classical”, we mean that the energy of a <strong>fluid</strong> particle depends only on local variables such as thedensity ρ and the temperature T . Van der Waals showed that it is actually impossible [Rowl<strong>in</strong>sonand Widom, 1982]: the <strong><strong>in</strong>terface</strong> would be sharp and surface tension would be null. That is whynon-local terms have to be considered. In the case of a liquid-vapor <strong><strong>in</strong>terface</strong>, van der Waalspostulated the follow<strong>in</strong>g thermodynamic description:F = F 0 (ρ, T ) + λ 2 (∇ρ)2 (14)where F is the volumetric free energy of the <strong>fluid</strong>, F 0 is its “classical” part and λ is the capillarycoefficient. For the sake of simplicity, we will always consider that λ is constant.F 0AρFigure 2: Illustration of the graph of the classical volumetric free energy F 0 (ρ).It can be shown that this particular form is justified from a molecular po<strong>in</strong>t of view. We willnot proove this and the <strong>in</strong>terested reader can refer to [Rocard, 1967] for <strong>in</strong>stance. In particular, itcan be shown that the value of λ depends only on the <strong>in</strong>termolecular potential.2.1.2 General equilibrium conditionsFor the sake of generality, we will consider that the volumetric free energy of the <strong>fluid</strong> is givenby the general expression F (ρ, T, ∇ρ). The differential of F thus readsdF = −S dT + g dρ + φ · d∇ρ (15)which def<strong>in</strong>es the entropy S, the Gibbs free enthalpy g as well as φ.The second law of thermodynamics states that a closed and isolated system at equilibrium issuch that its entropy is maximum. Mathematically, this reads∫δ [S + L 1 U(S, ρ, ∇ρ) + L 2 ρ] dV = 0 (16)V9


where δ represents the variation, V is the <strong>fluid</strong> doma<strong>in</strong> S is the volumetric entropy, U is the volumetric<strong>in</strong>ternal energy and where L 1 and L 2 are two constant Lagrange multipliers account<strong>in</strong>gfor the constra<strong>in</strong>ts of conservation respectively of the energy and mass of the system.S<strong>in</strong>ceU = F + S Tone hasThusNow∫V∫[(1 + L 1 T ) δ S + (L 1 g + L 2 ) δρ + L 1 φ · δ∇ρ] dV = 0Vφ · δ∇ρ = φ · ∇(δρ)= ∇ · (φ δρ) − (∇ · φ) δρ[(1 + L 1 T ) δ S + (L 1 (g − ∇ · φ) + L 2 ) δρ] dV = 0 (17)where we have used the condition (that is discussed <strong>in</strong> section 2.4)∫∫∇ · (φ δρ) dV = n · φ δρ dS = 0Vwhere ∂V is the boundary of the doma<strong>in</strong> and n its unit normal outwardly directed and wherewe have assumed that n · φ = 0 on ∂V .S<strong>in</strong>ce equation (17) must be satisfied for any variation δS and δρ, one has:∂VT = − 1 L 1g − ∇ · φ = − L 2L 1S<strong>in</strong>ce L 1 and L 2 are constants, the equilibrium conditions readT = cste (18)g − ∇ · φ = cste (19)The first condition means that the temperature of the system is uniform at equilibrium. Thesecond condition means that the generalized Gibbs free enthalpy˜g ˆ= g − ∇ · φ (20)is uniform at equilibrium. This latter condition is a generalization of the classical equilibriumcondition stat<strong>in</strong>g that the Gibbs free enthalpy is uniform at equilibrium.It must be emphasized that these equilibrium condition are valid <strong>in</strong> the entire two-phasesystem, <strong>in</strong>clud<strong>in</strong>g the bulk liquid and vapor phases as well as the <strong>in</strong>terfacial zones.2.1.3 Internal structure of the <strong><strong>in</strong>terface</strong>In this section, we discuss the consequences of the equilibrium conditions derived <strong>in</strong> the previoussection on the <strong>in</strong>ternal structure of a liquid-vapor <strong><strong>in</strong>terface</strong>. For that purpose, we restrictthe analysis to the case where the expression for F (ρ, T, ∇ρ) is given by (14). In this case, theequilibrium condition (19) reads∂F 0∂ρ (ρ, T 0) − λ ∇ 2 ρ = cste (21)10


where T 0 is a constant correspond<strong>in</strong>g to the equilibrium temperature of the system; T 0 can beviewed as a parameter and is dropped <strong>in</strong> the follow<strong>in</strong>g developments.Mathematically, this equation is a differential equation that the density field ρ(x) must satisfyat equilibrium. To better understand the consequences of this differential equation, we nowconsider a planar <strong><strong>in</strong>terface</strong> at equilibrium at we denote z the coord<strong>in</strong>ate normal to the <strong><strong>in</strong>terface</strong>.We thus seek for the function ρ(z) that def<strong>in</strong>es the profile of the <strong>in</strong>terfacial zone. This functionmust satisfywhere g 0 ˆ= ∂F 0 /∂ρ.g 0g 0 (ρ) − λ d2 ρ= cste (22)dz2 AρFigure 3: Illustration of the graph of g 0 (ρ).Very far from the <strong>in</strong>terfacial zone, bulk liquid and vapor phases exist and therefore d 2 ρ/dz 2 =0; likewise, dρ/dz = 0. This equation shows thatg 0 (ρ v ) = cste = g 0 (ρ l ) ˆ= g eq (23)where ρ v and ρ l are the densities of the vapor and liquid phases respectively far from the <strong><strong>in</strong>terface</strong>.This equation shows that the specific free Gibbs energy of the phases are equal. This correspondsto the condition of equilibrium of the <strong><strong>in</strong>terface</strong> discussed <strong>in</strong> section 1.2.2.By multiply<strong>in</strong>g equation (22) by dρ/dz and <strong>in</strong>tegrate, one getsF 0 (ρ) − F 0 (ρ v ) − g eq (ρ − ρ v ) = λ 2( ) 2 dρ(24)dzIf we def<strong>in</strong>eone haswhich is clearly a differential equation for ρ(z).W (ρ) ˆ= F 0 (ρ) − F 0 (ρ v ) − g eq (ρ − ρ v ) (25)λ2This equation shows <strong>in</strong> particular thatwhich is equivalent to( ) 2 dρ= W (ρ)dzW (ρ v ) = W (ρ l ) (26)F 0 (ρ l ) − g eq ρ l = F 0 (ρ v ) − g eq ρ v (27)Now, us<strong>in</strong>g classical thermodynamic relations, it can be shown that the pressure P 0 is given byP 0 = ρ g 0 − F 0 (28)11


Thus equation (27) readsP 0 (ρ v ) = P 0 (ρ l ) (29)This relation means that the liquid and vapor phases at equilibrium are equal. We recover theclassical condition of equilibrium of a planar <strong><strong>in</strong>terface</strong>.It is worth not<strong>in</strong>g that the two conditions (23) and (29) are two equations of the two unknownsρ v and ρ l that can thus be determ<strong>in</strong>ed. Moreover, these conditions have a simple graphical <strong>in</strong>terpretationon the graph of the function F 0 (ρ). Indeed, the condition (23) means that the two slopesof the tangent to this graph at ρ v and ρ l are equal. The condition (29) means that the y-<strong>in</strong>terceptsof these tangents are equal. Therefore, these two tangents are the same. Thus, ρ v and ρ l are def<strong>in</strong>edby the bi-tangent to the graph of the function F 0 (ρ). Moreover, this allows to show that W (ρ) (def<strong>in</strong>edby (25)) is actually the height between F 0 (ρ) and its bi-tangent.For the sake of simplicity (as will be shown hereafter), the function W (ρ) is often modeled asfollows:W (ρ) = A (ρ − ρ v ) 2 (ρ − ρ l ) 2 (30)where A is a parameter characteristic of the function F 0 (ρ) (cf. figure 2). This particular formallows to simplify the differential equation (24):This differential admits the follow<strong>in</strong>g solutionwheredρdz = √2 Aλ (ρ − ρ v) (ρ l − ρ)ρ(z) = ρ l + ρ v2where h represents the <strong><strong>in</strong>terface</strong> thickness.+ ρ l − ρ v2h ˆ= 1ρ l − ρ v√λ2 A( z)tanh2 h(31)2.1.4 Surface excess energyThis analysis also allows to determ<strong>in</strong>e the energy “concentrated” at the <strong><strong>in</strong>terface</strong>. This energy,denote F ex is def<strong>in</strong>ed as follows:whereF ex ˆ=∫ zi−∞(F (ρ, ∇ρ) − F 0 (ρ v )) dz +ρ ex ˆ=∫ zi−∞∫ +∞z i(ρ(z) − ρ v ) dz +(F 0 (ρ l ) − F (ρ, ∇ρ)) dz − g eq ρ ex∫ +∞z i(ρ l − ρ(z)) dzwhere z i is any position. Actually it is straightforward to show that F ex does not depend on z i .Us<strong>in</strong>g the relations (25) and (24), it can be shown thatF ex =∫ +∞−∞λ( ) 2 ∫ dρρl√ √dz = 2 λ W (ρ) dρ (32)dzρ vThis energy concentrated at the <strong><strong>in</strong>terface</strong> is <strong>in</strong>terpreted as the surface tension.It is worth not<strong>in</strong>g that the above expression is general and is valid for any expression for thefunction F 0 (ρ) and therefore W (ρ). In the particular case where W (ρ) is given by (30), one f<strong>in</strong>dsthat the expression for the surface tension is the follow<strong>in</strong>g:σ = (ρ l − ρ v ) 36√2 A λ (33)12


2.1.5 Equilibrium of a pherical <strong>in</strong>clusionIn section 1.2, we showed that surface tension has an effect on the outer bulk phases through theLaplace relation. Is it recovered by the van der Waals model?The equilibrium condition (21) is always valid and is <strong>in</strong> particular valid for a spherical <strong>in</strong>clusion(bubble or droplet) at equilibrium. In this case, we consider a spherical system of coord<strong>in</strong>ateswhose orig<strong>in</strong> is the center of the <strong>in</strong>clusion. The only variations that must accounted for are alongthe radial direction. The equilibrium condition thus readswhere r is the radial coord<strong>in</strong>ate.This equation shows <strong>in</strong> particular thatg − λ d2 ρdr 2 − λ 2 rdρ= cste (34)drg(ρ s v) = g(ρ s l ) ˆ= g s eq (35)This relation shows that the bulk phase specific Gibbs free enthalpies are equal. It is importantto note that, at this po<strong>in</strong>t, the value of geq s is unknown and is a priori different from g eq (theequilibrium value of a planar <strong><strong>in</strong>terface</strong>). This implies <strong>in</strong> particular that the densities of the bulkphases (ρ s v and ρs l) are different from those of a planar <strong><strong>in</strong>terface</strong>.By multiply<strong>in</strong>g equation (34) by dρ/dr and <strong>in</strong>tegrat<strong>in</strong>g from r = 0 to r yieldsF 0 (ρ) − F 0 (ρ s i ) − g s eq(ρ − ρ s i ) = λ 2( ) 2 ∫ dρr( ) 22 dρ+dr 0 η λ dη (36)dηwhere the subscript i denotes the <strong>in</strong>terior phase (i.e. vapor for a bubble or liquid for a drop).This expression shows <strong>in</strong> particular thatF 0 (ρ s e ) − F 0 (ρ s i ) − gs eq (ρs e − ρs i ) = ∫ ∞where the subscript e denotes the exterior phase.Us<strong>in</strong>g the def<strong>in</strong>ition (28) of the pressure, this relation reads:P 0 si∫ ∞( ) 2− Pe 0 s 2 dρ=0 η λ dηdη0( ) 22 dρη λ dηdηThe variations of ρ are significant only <strong>in</strong> the vic<strong>in</strong>ity of the radius R of the <strong>in</strong>clusion; thus∫ ∞0( ) 22 dρη λ dη ≃ 2 dη R∫ ∞0λ( ) 2 dρdηdηMoreover, if the density profile (determ<strong>in</strong>ed by the <strong>in</strong>tegro-differential equation (36)) is onlyweakly <strong>in</strong>fluenced by the curvature effects the last <strong>in</strong>tegral term can be approximated by tak<strong>in</strong>gthe density profile of a planar <strong><strong>in</strong>terface</strong>, <strong>in</strong> which caseThereforeThis is the Laplace relation.∫ ∞0λP 0 si( ) 2 dρdη ≃ σdη− P 0 se≃ 2 σRThe graphical determ<strong>in</strong>ation of the condition of equilibrium of a spherical <strong>in</strong>clusion is illustrated<strong>in</strong> figure 4.13


WW2σ/Rρρ2σ/RbubbledropFigure 4: Graphical determ<strong>in</strong>ation of the states of a spherical <strong>in</strong>clusion at equilibrium.2.1.6 Expression for the volumetric free energyThe study of of the equilibrium conditions developed <strong>in</strong> the previous sections shows the importanceof the dependence <strong>in</strong> ρ of the volumetric free energy F 0 on the determ<strong>in</strong>ation of the <strong>in</strong>ternalstructure of the <strong><strong>in</strong>terface</strong> and on the value of the surface tension. At a given temperature T 0 , theexpression for F 0 (ρ) is the follow<strong>in</strong>g (cf. relations (25) and (28)):F 0 (ρ) = W (ρ) + g eq ρ − P eqwhere P eq is the saturation pressure at the temperature considered. The equilibrium conditionsstudied <strong>in</strong> the previous sections show that the values of P eq and of g eq have no <strong>in</strong>fluence the resultsfor an isothermal system. Therefore, very often, one takes F 0 (ρ) = W (ρ).Because the equilibrium condition on the temperature is simply T = cste, <strong>in</strong> the previoussections, the temperature of the system has been considered as a parameter. However, the effectsof the temperature are important to account for accurately. Indeed, <strong>in</strong> many applications, phasechangeoccurs because the system is heated, which means that the temperature gradients drivethe phase transition. Moreover, the latent heat is also an important thermodynamic propertythat has to be accounted for. The general expression (15) for the differential of F shows that itsdependence <strong>in</strong> T is related to the entropy:s = − 1 ( ) ∂Fρ ∂TρIn particular, the latent heat L is such that[ ( )1 ∂FL = T (s v − s l ) = T(ρ v (T )) − 1 ( ) ]∂F(ρ l (T ))ρ v (T ) ∂Tρ l (T ) ∂Twhere ρ v (T ) and ρ l (T ) are the vapor and liquid densities at saturation.Moreover, the heat capacity at constant volume Cv (another important thermo-physical property)is given by( ) ∂sCv = T(38)∂TρThus the dependence <strong>in</strong> T of F must be consistent with the data L(T ) and Cv(T ). Without anyapproximation, it is difficult to provide an expression for F 0 (ρ, T ) that satisfies this consistency.To keep the simplicity of the polynomial expression (30) for W (ρ), we can make the follow<strong>in</strong>gapproximation:F 0 (ρ, T ) = W (ρ, T ) + g eq (T ) ρ − P eq (T )(37)14


which is the equation of evolution of the <strong>in</strong>ternal energy.The expression for the <strong>in</strong>ternal energy isu = F ρ + T s (44)and the pressure P is def<strong>in</strong>ed byso that the differential of u reads( ) ∂FP ˆ= ρ − F (45)∂ρρ du = ρ T ds + P ρdρ + φ · d∇ρ (46)Us<strong>in</strong>g equations (39), (42) and (43) to express dρ/dt, ds/dt and du/dt <strong>in</strong> the above relation, onehas−∇ · q + T : ∇v = T (−∇ · q s + ∆ s ) − P ∇ · v + φ · d∇ρdtThe last term of this equation is transformed as follows(47)φ · d∇ρdt( )∂∇ρ= φ · + v · ∇∇ρ∂t( ) ∂ρ= φ · ∇ + φ∂t i v j ρ ,ij(= ∇ · φ ∂ρ )∂t(= ∇ · φ ∂ρ )∂t(= ∇ · φ ∂ρ∂t− ∂ρ∂t (∇ · φ) + (φ i v j ρ ,j ) ,i− (φ i v j ) ,iρ ,j− ∂ρ∂t (∇ · φ) + ∇ · (φ v · ∇ρ) − φ i,i v j ρ ,j − φ i v j,i ρ ,j)− ∂ρ (∇ · φ) + ∇ · (φ v · ∇ρ) − (v · ∇ρ) ∇ · φ − (φ ⊗ ∇ρ) : ∇v∂twhere ψ ,i ≡ ∂ψ/∂x i and where the E<strong>in</strong>ste<strong>in</strong> convention on the repeated <strong>in</strong>dices has been used.Thusφ · d∇ρ (= ∇ · φ dρ )− dρ (∇ · φ) − (φ ⊗ ∇ρ) : ∇vdtdt dtSubstitut<strong>in</strong>g this relation <strong>in</strong> equation (47) allows to express the entropy source ∆ s as follows[∆ s = ∇ · q s − 1 (q + φ dρ )]− 1 (T dt T 2 q + φ dρ )· ∇T + 1 [T + (P − ρ ∇ · φ) I + φ ⊗ ∇ρ] : ∇vdt T(48)The second law of thermodynamics states that ∆ s ≥ 0 for any motion. The follow<strong>in</strong>g expressionsfor q and T satisfy this condition 3 :3 These relations are not the most general. For <strong>in</strong>stance, the thermal conductivity k is <strong>in</strong> general a tensor of order twoof the form k = k 1 I + k 2 ∇ρ ⊗ ∇ρ/(∇ρ) 2 . This relation expresses that the thermal conductivity <strong>in</strong> the normal andtangential directions to the <strong><strong>in</strong>terface</strong> can be different. Likewise, a Newtonian behavior is very restrictive compared to thegeneral expressions found for the dissipative stress tensor <strong>in</strong> which five different “viscosity” coefficients appear.It is worth not<strong>in</strong>g that, despite its simplicity, the method used to derive these expressions, and especially the expressionfor q, might not be the most rigorous from a fundamental po<strong>in</strong>t of view. Indeed, us<strong>in</strong>g the Hamilton’s pr<strong>in</strong>cipal, it canbe shown that the term φ dρ/dt is actually not a heat flux per say but is rather a work s<strong>in</strong>ce it has no contribution tothe entropy source and is thus a conservative contribution. This term is known as the “<strong>in</strong>tersticial work<strong>in</strong>g” [Dunn andSerr<strong>in</strong>, 1965].16


q = −φ dρ − k ∇T (49)dtT = (−P + ρ ∇ · φ) I − φ ⊗ ∇ρ + τ (50)where k must be positive and where τ is the dissipative stress tensor that must satisfyτ : ∇v ≥ 0A classical Newtonian <strong>fluid</strong> satisfies this condition.Us<strong>in</strong>g these closure relations, the system of balance equations that describes the motion of the<strong>fluid</strong> is the follow<strong>in</strong>g:dρ= −ρ ∇ · v (51)dtρ dvdt= −∇P + ∇(ρ ∇ · φ) − ∇ · (φ ⊗ ∇ρ) + ∇ · τ (52)ρ de (dt = ∇ · φ dρdt2.2.1 Korteweg stress tensor and surface tension force)+ ∇ · (k ∇T ) + ∇ · (v · T ) (53)In the most classical case where the energy of the <strong>fluid</strong> is expressed by (14),φ = λ ∇ρandP = P 0 (ρ, T ) − λ 2 (∇ρ)2The momentum balance equation thus readsρ dvdt = −∇P 0 + ∇(λ ρ ∇ 2 ρ + λ )2 (∇ρ)2 − ∇ · (λ∇ρ ⊗ ∇ρ) + ∇ · τ (54)The stress tensor (λ∇ρ ⊗ ∇ρ) is called the Korteweg stress tensor [Korteweg, 1901]. We show<strong>in</strong> the follow<strong>in</strong>g that this stress tensor implies a tension force <strong>in</strong> the tangential direction to the<strong><strong>in</strong>terface</strong>, <strong>in</strong>terpreted as the surface tension force.To simplify the analysis, we consider a planar <strong><strong>in</strong>terface</strong> at equilibrium. We denote z and x thecoord<strong>in</strong>ates respectively normal and tangential to the <strong><strong>in</strong>terface</strong> (because of symmetry, only onetangential direction can be accounted for). Thus, all the variables depend only on z.Let us first def<strong>in</strong>e˜P ˆ= P 0 −(λ ρ ∇ 2 ρ + λ )2 (∇ρ)2so that, at equilibrium, the momentum balance equation (55) reduces toBy <strong>in</strong>tegration, one simply getsd ˜Pdz + λ d dz˜P (z) = P ∞ − λwhere P ∞ <strong>in</strong> the pressure <strong>in</strong> the bulk phases.( ) 2 dρ= 0dz( ) 2 dρdz17


S<strong>in</strong>ce λ is constant, s = s 0 . Therefore( ) ∂sρ T ds = ρ Cv 0 0dT + ρ T∂ρTdρThusρ Cv 0 dT( ) ∂s0dt = ∇ · (k ∇T ) − ρ T ∂ρTdρ+ τ : ∇v (55)dtThis equation is very similar to the classical equation of evolution of the temperature <strong>in</strong> as<strong>in</strong>gle-phase <strong>fluid</strong>. However, the <strong>in</strong>terpretation of the term <strong>in</strong> dρ/dt is very different. Indeed, for as<strong>in</strong>gle-phase <strong>in</strong>compressible <strong>fluid</strong>, this term vanishes (by def<strong>in</strong>ition). In the case of a liquid-vapor<strong>in</strong>terfacial zone, (dρ/dt) represents the rate of vaporization denoted γ c . Indeed, let us consideran <strong><strong>in</strong>terface</strong> where vaporization occurs. If we follow a liquid <strong>fluid</strong> particle <strong>in</strong> its motion, as itcrosses the <strong><strong>in</strong>terface</strong> to become vapor, its density drastically decreases, therefore dρ/dt < 0. Now,ρ T (∂s 0 /∂ρ) T is analogous to the latent heat (cf. (37)). Therefore, this term can be approximatedby γ c L ρ/(ρ l − ρ v ). This term is thus a spread<strong>in</strong>g of the latent heat source over the <strong>in</strong>terfacialzone.2.3 Different forms of the momentum balance equationIn the form (52) of the momemtum balance equation, the Korteweg tensor might not be the mostappropriate. Indeed, from a numerical po<strong>in</strong>t of view for <strong>in</strong>stance, the discretization of this termis not straightforward. Moreover, we will see that there exists an equivalent from <strong>in</strong> which onlyLaplacian and gradient operators appear, which is generally much easier to implement.The follow<strong>in</strong>g identity holds:∇ · (φ ⊗ ∇ρ) = ∇(ρ ∇ · φ) − ρ ∇(∇ · φ) + φ · ∇∇ρGiven the expression (45) for P and the differential of F (15), one hasThus, equation (52) readsρ dvdt∇P = ρ ∇g + ρ s ∇T − φ · ∇∇ρ= −ρ ∇ (g − ∇ · φ) − ρ s ∇T + ∇ · τThis form of the momemtum balance equation makes clearly appears the two conditions ofequilibrium (18)-(19). This shows <strong>in</strong> particular that if any of these thermodynamic equilibriumconditions is not satisfied, it triggers a <strong>fluid</strong> motion. Moreover, s<strong>in</strong>ce the thermodynamic equilibriumconditions appear <strong>in</strong> this equation, it shows its thermodynamic consistency. Moreover, froma numerical po<strong>in</strong>t of view, Jamet et al. [2002] showed that, through a detailed analysis of the discretizedenergy exchanges, this form allows to get rid of the so-called parasitic currents [Brackbillet al., 1992]. These non-physical currents, concentrated <strong>in</strong> the close vic<strong>in</strong>ity to the <strong><strong>in</strong>terface</strong>, are<strong>in</strong>duced by numerical truncation that are very difficult to elim<strong>in</strong>ate without a detailed analysisof the energy exchanges. The thermodynamic consistency of this model gives a framework todevelop accurate numerical schemes.2.4 Boundary conditions and contact angleSo far, <strong>in</strong> the determ<strong>in</strong>ation of the equilibrium conditions or of the equations of motion, we havenot studied the boundary conditions that must be applied on the density field. In this section,19


we show that particular boundary conditions arise from the existence of the capillary term <strong>in</strong> theenergy functional and that these boundary conditions are related to the contact angle 4 .Let us consider a liquid-vapor system <strong>in</strong> contact with a solid wall. Let us consider an energyof <strong>in</strong>teraction between the solid and the <strong>fluid</strong> U s (per unit surface area) and let us assume thatthis energy depend only on the local density of the <strong>fluid</strong> at the boundary (this assumption can bejustified by a mean-field approximation [Gou<strong>in</strong>, 1998]). Therefore, the total <strong>in</strong>ternal energy of thesystem is∫∫U(S, ρ, ∇ρ) dV + U s (ρ) dAV∂Vwhere ∂V is the boundary of the <strong>fluid</strong> doma<strong>in</strong> V . Follow<strong>in</strong>g the developments made <strong>in</strong> section2.1.2, the application of the second law of thermodynamics to determ<strong>in</strong>e the equilibriumconditions of the system yields∫∫δ [S + L 1 U(S, ρ, ∇ρ) + L 2 ρ] dV + δ L 1 U s (ρ) dA = 0Vwhere we rem<strong>in</strong>d that L 1 is the constant Lagrange multiplier account<strong>in</strong>g for the constra<strong>in</strong>t ofconservation of the total <strong>in</strong>ternal energy. This yields (cf. section 2.1.2):∫∫ ( )dUs[(1 + L 1 T ) δ S + (L 1 (g − ∇ · φ) + L 2 ) δρ] dV +dρ + n · φ δρ dS = 0VThe last surface <strong>in</strong>tegral is a term that did not appear <strong>in</strong> the study developed <strong>in</strong> section 2.1.2.S<strong>in</strong>ce the above condition must be satisfied for any variation δρ, the follow<strong>in</strong>g condition must besatisfied at the boundary ∂V :n · φ = − dU s(56)dρTo illustrate the physical mean<strong>in</strong>g of this boundary condition, let us consider the follow<strong>in</strong>g assumptions:the expression for F (ρ, ∇ρ, T ) is given by (14) so that φ = λ ∇ρ and U s (ρ) is assumedto be l<strong>in</strong>ear so that dU s /dρ = β = cste. The boundary condition therefore reads∂V∂Vn · ∇ρ = − β λ(57)S<strong>in</strong>ce β and λ are constant, this consdition imposes the value of the normal derivative of ρ. Asillustrated <strong>in</strong> figure 6, this imposes the value of the contact angle. The <strong>in</strong>terested reader can referto [Seppecher, 1996, Jacqm<strong>in</strong>, 2000] for <strong>in</strong>stance where this boundary condition has been studied.It is worth not<strong>in</strong>g that the boundary condition (56) is an equilibrium boundary condition.However, this condition can be extended to out-of-equilibrium conditions to recover a variationof the contact angle with the speed of displacement of the contact l<strong>in</strong>e (a variation that is observedexperimentally). This form is thermodynamically coherent s<strong>in</strong>ce it ensures that the entropy of thesystem <strong>in</strong>creases.3 Two-phase flows of non-miscible <strong>fluid</strong>s: the Cahn-Hilliard modelIn the previous section, we presented the van der Waals model of capillarity dedicated to themodel<strong>in</strong>g of an <strong><strong>in</strong>terface</strong> that separates the liquid and vapor phase of the same pure substance.Many two-phase systems are made of different substances, for <strong>in</strong>stance air and water or oil andwater. In this case, an <strong><strong>in</strong>terface</strong> separates two phases that are made of different species. Nonmisciblesphases are phases for which no mass transfer from one phase to the other exist.4 When a drop of liquid is put <strong>in</strong> contact with a solid wall, it is observed that the angle θ between the liquid-gas<strong><strong>in</strong>terface</strong> and the solid surface is characteristic of the triplet solid-liquid-gas. For <strong>in</strong>stance, for the same liquid and gas,if we change the nature of the solid, the same volume of liquid either tends to spread on the surface (θ < 90 ◦ ) or toretract (θ > 90 ◦ ). This tendency is related to the energy of <strong>in</strong>teraction between the solid and the liquid. If the solid hasmore aff<strong>in</strong>ity with the liquid than with the gas, the system tends to m<strong>in</strong>imize its energy by <strong>in</strong>creas<strong>in</strong>g the area of contactbetween the liquid and the solid (θ < 90 ◦ ).20


vaporliquid¤¡¤¡¤¡¤ §¡§¡§¡§ ¨¡¨¡¨¡¨ ©¡©¡©¡© ¡¡¡ ¡¡¡ ¡¡¡ ¡¡¡ ¡¡¡ £¡£¡£¡£¤¡¤¡¤¡¤ §¡§¡§¡§ ¨¡¨¡¨¡¨ ©¡©¡©¡© ¡¡¡ ¡¡¡ ¡¡¡ ¡¡¡ ¡¡¡ £¡£¡£¡£¤¡¤¡¤¡¤ §¡§¡§¡§ ¨¡¨¡¨¡¨ ©¡©¡©¡© ¡¡¡ ¡¡¡ ¡¡¡ ¡¡¡ ¡¡¡ £¡£¡£¡£¦¡¦¡¦ ¥¡¥¡¥¤¡¤¡¤¡¤ §¡§¡§¡§ ¨¡¨¡¨¡¨ ©¡©¡©¡© ¡¡¡ ¡¡¡ ¡¡¡ ¡¡¡ ¡¡¡ £¡£¡£¡£¦¡¦¡¦ ¥¡¥¡¥θ∇ρ¤¡¤¡¤¡¤¥¡¥¡¥ ¦¡¦¡¦§¡§¡§¡§ ¨¡¨¡¨¡¨ ©¡©¡©¡© ¡¡¡ ¡¡¡ ¡¡¡ ¡¡¡ ¡¡¡ £¡£¡£¡£solidn¤¡¤¡¤¡¤ §¡§¡§¡§ ¨¡¨¡¨¡¨ ©¡©¡©¡© ¡¡¡ ¡¡¡ ¡¡¡ ¡¡¡ ¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡£¡£¡£¡£¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡¡Figure 6: Illustration of the contact angle boundary condition.3.1 Thermodynamic modelFrom a physical po<strong>in</strong>t of view, why do these two species tend to be separated? At the molecularlevel, the energy of <strong>in</strong>teraction of two molecules of different species is larger than that one twomolecules of the same species. Thus, s<strong>in</strong>ce the system tends to m<strong>in</strong>imize its energy, if the speciesare separated, the energy of the system is weaker. This simple reason<strong>in</strong>g shows that the “driv<strong>in</strong>gforce” of the transition, which eventually gives rise to the existence of an <strong><strong>in</strong>terface</strong>, is the amountof one species <strong>in</strong>to the other. If not too many molecules of one species is present is the other,the energy of the system is not that <strong>in</strong>creased but if more molecules are added, the energy getsto large and the system tends to separate <strong>in</strong>to two different phases: one rich <strong>in</strong> the first speciesand the other rich <strong>in</strong> the other species. In this case, the relevant thermodynamic variable is theconcentration of one species <strong>in</strong> the mixture.3.1.1 A mean-field approximationFor the sake of simplicity, we will first consider that the density of the mixture is constant forany value of the concentration of one species <strong>in</strong> the mixture; it is denoted ρ 0 . Let c denote themass fraction (or concentration) of one species <strong>in</strong> the mixture. By analogy with the van der Waalsmodel, Cahn and Hilliard [Cahn and Hilliard, 1958, 1959a,b] postulated that the free energy ofthe system F is given byF = F 0 (c) + λ 2 (∇c)2 (58)where F 0 (c) is the “classical” part of the energy and λ is the capillary coefficient.F 0Aµ 0 AccFigure 7: Illustration of the graph of the functions F 0 (c) and µ 0 (c).However, it can be shown that this particular form is justified from a molecular po<strong>in</strong>t of view.Indeed, us<strong>in</strong>g a mean-field approximation, it can be shown that the attractive energy of <strong>in</strong>teractionof molecules of different types gives rise to this form for the energy of the mixture and thatλ depends only on the <strong>in</strong>ter-molecular potentials.21


3.1.2 Equilibrium conditionsS<strong>in</strong>ce we deal with a mixture, we can assume that thermal effects are negligible. We can thusassume that the temperature of the system is imposed to a constant. S<strong>in</strong>ce the system is assumedto be isothermal, the equilibrium is characterized by a m<strong>in</strong>imum of a m<strong>in</strong>imum of its free energy.Mathematically, this reads: ∫δ (F (c, ∇c) + L ρ 0 c) dV = 0Vwhere L is a constant Lagrange multiplier account<strong>in</strong>g for the fact that the system is closed andthat the total mass of each species is constant, which reads∫ρ 0 c dV = csteVFollow<strong>in</strong>g the same developments as those presented <strong>in</strong> section 2.1.2, one f<strong>in</strong>ds that the equilibriumcondition is the follow<strong>in</strong>g:( )∂F ∂F∂c − ∇ · = cste∂∇cThis condition is very similar to the condition (19) found for the van der Waals model.In the particular case where F (c, ∇c) is given by the expression (58), this equilibrium conditionsimply readswhereµ 0 (c) − λ∇ 2 c = cste (59)µ 0 (c) ˆ= dF 0(60)dcThis equation is the same as that obta<strong>in</strong>ed for the van der Waals model and the same conclusionshold.In particular, for a planar <strong><strong>in</strong>terface</strong> at equilibrium, it is found that the equilibrium conditionscorrespond to the double-tangent to the graph of the function F 0 (c). This condition def<strong>in</strong>es themass fraction of the phases at equilibrium of a planar <strong><strong>in</strong>terface</strong>, c 1 and c 2 , as well as the chemicalpotential of a planar <strong><strong>in</strong>terface</strong> µ eq . It is worth emphasiz<strong>in</strong>g that c 1 and c 2 are not equal to 0or 1; actually, from a physical po<strong>in</strong>t of view c 1 and c 2 cannot be exactly equal to 0 or 1. Theirvalue actually depend on the physical system considered. It is then convenient to <strong>in</strong>troduce thedouble-well function W (c) def<strong>in</strong>ed as the difference between F (c) and its double-tangent:W (c) ˆ= F (c) − (F (c 1 ) + µ eq (c − c 1 ))This function is very often approximated by a polynomial of degree 4:W (c) = A (c − c 1 ) 2 (c − c 2 ) 2It is worth not<strong>in</strong>g that this particular form can be justified from a mean-field approximation closeto a critical po<strong>in</strong>t. It must be emphasized that this approximation is valid when c 1 ≃ c 2 , whichmeans <strong>in</strong> particular that it is not valid for c 1 ≃ 0 and c 2 ≃ 1. This is important because, often,Cahn-Hilliard <strong>models</strong> are used with c 1 = 0 and c 2 = 1, which actually only corresponds to arenormalization of the “true” mass fraction.Moreover, the study of the equilibrium of a spherical <strong>in</strong>clusion implies that same results as <strong>in</strong>the van der Waals model: (i) the chemical potentials of the bulk phases surround<strong>in</strong>g the curved<strong><strong>in</strong>terface</strong> are equalµ 0 (c i ) = µ 0 (c e ) ˆ= µ s e22


and (ii) there is an equivalent of the Laplace relationwhere the “pressure” P 0 is def<strong>in</strong>ed byP 0 (c i ) − P 0 (c i ) = 2 σRP 0 (c) ˆ= c dF 0dc − F 0 (c)In this case, the “pressure” P 0 does not have the <strong>in</strong>terpretation of a physical pressure.A Taylor expansion of these equilibrium conditions of a spherical <strong>in</strong>clusion around the equilibriumstate of a planar <strong><strong>in</strong>terface</strong> allow to show thatµ s e ≃ µ eq + 1 2 σc i − c e R(61)It is worth not<strong>in</strong>g that the form of this equilibrium condition is actually similar to the Gibbs-Thompson condition (12).This equilibrium condition shows <strong>in</strong> particular that the value of the chemical potential is proportionalto the curvature of the <strong><strong>in</strong>terface</strong>. This expla<strong>in</strong>s how an <strong><strong>in</strong>terface</strong> tends to get spherical.Indeed, let us consider an closed <strong><strong>in</strong>terface</strong> whose shape is <strong>in</strong>itially irregular. If, at each po<strong>in</strong>t ofthe <strong><strong>in</strong>terface</strong>, the <strong>in</strong>terfacial zone is at local thermodynamic equilibrium, the above equilibriumcondition is satisfied locally. These means <strong>in</strong> particular that, along the <strong><strong>in</strong>terface</strong>, the chemicalpotential is not uniform. Accord<strong>in</strong>g to the Cahn-Hilliard equation, this yields a diffusion massflux and therefore a mass diffusion. This mass diffusion makes the overall system evolve and,s<strong>in</strong>ce the Cahn-Hilliard equation is thermodynamically coherent, this evolution tends to makethe system get closer to an equilibrium state. If we consider a spherical <strong>in</strong>clusion at equilibrium,the chemical potential along the <strong><strong>in</strong>terface</strong> is constant, therefore no mass flux exist and the systemkeeps at rest.3.2 The Cahn-Hilliard equationIn the previous section, we derived the equilibrium condition for a Cahn-Hilliard <strong>fluid</strong>. The issueis to determ<strong>in</strong>e how the system behaves out of equilibrium. To start this analysis, we will firstsimplify the system by assum<strong>in</strong>g that the velocity of the mixture is null. In this case, the evolutionof the concentration evolves through a mass diffusion equation that reads∂c∂t = −∇ · jwhere j is the diffusion mass flux, whose expression must be determ<strong>in</strong>ed.The evolution of the free energy of the mixture is given by∂F∂t = −∇ · q − ∆ fwhere q is the heat flux and ∆ f is the energy dissipation. The second law of thermodynamicsimposes that ∆ f ≥ 0.Us<strong>in</strong>g the same developments as those presented <strong>in</strong> section 2.2 for the van der Waals model,one f<strong>in</strong>ds that[∆ f = −j · ∇ (µ − ∇ · φ) − ∇ · q + φ ∂c]∂t − j (µ − ∇ · φ)whereµ ˆ= ∂F∂c23


φ ˆ= ∂F∂∇cThe follow<strong>in</strong>g expression for j ensures that ∆ f ≥ 0j = −κ ∇ (µ − ∇ · φ)where κ is called the mobility. It is worth not<strong>in</strong>g that κ ≥ 0 and is any function of the thermostaticparameter of the system (i.e. c and ∇c) 5 .In the particular case where F (c, ∇c) is given by (58), the mass diffusion equation reads∂c∂t = ∇ · [κ∇ ( µ 0 (c) − λ∇ 2 c )] (62)This equation is called the Cahn-Hilliard equation. It is worth not<strong>in</strong>g that it is a generalizedmass diffusion equation that is valid <strong>in</strong> the entire two-phase system.The Cahn-Hilliard equation is a generalized mass diffusion equation. In particular, it shows thatthe local diffusion mass flux is proportional to the gradient of the generalized chemical potential(µ 0 (c) − λ∇ 2 c).3.3 Equations of motionIn the previous section, we derived the Cahn-Hilliard equation. This equation is the only equationthat has to be accounted for if the velocity of the mixture is null. In this section, we derivethe equations of motion of the mixture when this velocity is not null.3.3.1 Bouss<strong>in</strong>esq approximationIn this section, we assume that the density of the mixture does not depend on the compositionof the mixture: ρ(c) = ρ 0 . This the density is constant, the general mass balance equation of themixture∂ρ+ ∇ · (ρ v) = 0∂tdegenerates to the solenoidal condition∇ · v = 0In this case, we know that, <strong>in</strong> the momemtum balance equation, the pressure is the Lagrangemultiplier associated to this solenoidal constra<strong>in</strong>t. Us<strong>in</strong>g a similar approach as that developed <strong>in</strong>section 2.2 for the van der Waals model, it can be shown that the momemtum balance equationreadsρ 0∂v∂t + ρ 0 v · ∇v = −∇P − ∇ · (λ∇c ⊗ ∇c) + ∇ · τwhere τ can be expressed as the classical viscous stress tensor.The Cahn-Hilliard equation (62) has to be modified to account for the convection term asfollows:∂c∂t + v · ∇c = ∇ · [κ∇ ( µ 0 (c) − λ∇ 2 c )]5 In the general case, the mobility is a tensor of the formκ = κ 1 I + κ 2 ∇c ⊗ ∇c/(∇c) 2andj = −κ · (µ − ∇ · φ)24


It can be shown that this system of equations is thermodynamically consistent <strong>in</strong> the sensethat the total energy of the system is a decreas<strong>in</strong>g function of time:∫ddt V[F 0 (c) + λ 2 (∇c)2 + ρ 0 v 22] ∫dV = −{κ [ ∇ ( µ 0 (c) − λ∇ 2 c )] }2+ τ : ∇v dV < 0VTo account for buoyancy effects the gravity term ρ g (where g is the acceleration of gravity)has to be added <strong>in</strong> the momentum balance equation. If ρ = ρ 0 <strong>in</strong> this force, this force has noeffect on the flow: it only modifies the pressure (the pressure P can be replaced by the pressure(P − ρ 0 g z), which does not modify the structure of the equations). For the gravity to have aneffect on the flow and <strong>in</strong> particular to account for a buoyancy effect, variations of the density mustbe accounted for. However, these variations are a priori not compatible with the <strong>in</strong>compressibilityapproximation. That is why the Bouss<strong>in</strong>esq approximation is generally used: the variation of thedensity is neglected except <strong>in</strong> the gravity term where it is l<strong>in</strong>earized. The momentum balanceequation therefore readswhereρ 0∂v∂t + ρ 0 v · ∇v = −∇P − ∇ · (λ∇c ⊗ ∇c) + ∇ · τ + ρ 0 (1 + β(c − c 0 )) gWe summarize the system of equations:β ˆ= 1 ( ) dρρ 0 dc0∇ · v = 0 (63)∂c+ v · ∇c = ∇ · [κ ∇˜µ] (64)∂t˜µ ˆ= µ 0 (c) − λ∇ 2 c (65)∂vρ 0∂t + ρ 0 v · ∇v = −∇P − ∇ · (λ∇c ⊗ ∇c) + ∇ · τ + ρ 0 (1 + β(c − c 0 )) g (66)(µ 0 (c) = 4 A (c − c 1 ) (c − c 2 ) c − c )1 + c 2(67)2It is worth not<strong>in</strong>g that, like <strong>in</strong> the van der Waals model (cf. section 2.3), the “stress form” of themomentum balance equation can be transformed <strong>in</strong>to an equivalent “potential form” <strong>in</strong> whichthe generalized chemical potential appears [Jacqm<strong>in</strong>, 1999]:ρ 0∂v∂t + ρ 0 v · ∇v = −∇ ˜P + ˜µ ∇c + ∇ · τ + ρ 0 (1 + β(c − c 0 )) g (68)˜P ˆ= P + FThis form of the momentum balance equation shows the <strong>in</strong>fluence of the curvature on themomemtum balance equation. Indeed, we showed <strong>in</strong> the previous section that, at equilibriumof a spherical <strong>in</strong>clusion, the value of the chemical potential is proportional to the local <strong><strong>in</strong>terface</strong>curvature (cf. (61)). Therefore, the term ˜µ ∇c represents a spread<strong>in</strong>g (over the <strong><strong>in</strong>terface</strong> thickness)of a force proportional to the <strong><strong>in</strong>terface</strong> curvature and oriented <strong>in</strong> the direction normal to the <strong><strong>in</strong>terface</strong>approximated by ∇c. This <strong>in</strong>terpretation is very close to the Cont<strong>in</strong>uous Surface Forcecommonly used <strong>in</strong> sharp <strong><strong>in</strong>terface</strong> numerical methods [Brackbill et al., 1992].This system of equations is particularly attractive numerically. Indeed, we one has a numericalcode dedicated to the simulation of <strong>in</strong>compressible flows, it can be very easily generalized totwo-phase capillary flows. Indeed, one only needs to implement a source term <strong>in</strong> the momentum25


alance equation and a convection-diffusion-like equation for c (the Cahn-Hilliard equation). Anexample is given <strong>in</strong> figure 8. The system simulated is the impact of a heavier droplet on a solidgrid. This figure shows <strong>in</strong> particular that topological changes are automatically accounted for.Us<strong>in</strong>g standard second order schemes for the discretization <strong>in</strong> space, the <strong><strong>in</strong>terface</strong> is captured byabout 4 mesh cells. However, Jacqm<strong>in</strong> [1999] shows that it is possible to reduce it to about 2 byus<strong>in</strong>g a more complex scheme.Figure 8: Numerical simulation of the impact of a droplet on a grid us<strong>in</strong>g the Cahn-Hilliardmodel with a Bouss<strong>in</strong>esq approximation.Because of its simplicity, simulat<strong>in</strong>g three dimensional systems is particularly easy, even onparallel computers. An illustration of the complex three dimensional systems that can be simulatedis shown <strong>in</strong> figure 9.Figure 9: Numerical simulation of a complex Rayleigh-Taylor <strong>in</strong>stability <strong>in</strong> a I-shape reservoirus<strong>in</strong>g a Cahn-Hilliard diffuse <strong><strong>in</strong>terface</strong> model.3.3.2 F<strong>in</strong>ite density contrastThe case where the bulk phases have very different densities (for <strong>in</strong>stance air and water at roomtemperature ρ a ≃ 1 kg/m 3 and ρ w ≃ 1000 kg/m 3 ) is rather different. From a physical po<strong>in</strong>t of26


view, <strong>in</strong> this case, the velocity is not solenoidal [Galdi et al., 1991]. Therefore, a priori, the pressurecan no longer be <strong>in</strong>troduced as a Lagrange multiplier account<strong>in</strong>g for this constra<strong>in</strong>t. However,from a practical po<strong>in</strong>t of view, the solenoidal constra<strong>in</strong>t on the velocity field is attractive: it allowsto use standard numerical methods such as the projection method. The natural choice is thus touse the follow<strong>in</strong>g equations∇ · v = 0 (69)dc= ∇ · [κ ∇˜µ]dt(70)ρ(c) dvdt = −∇ ˜P + ˜µ ∇c + ∇ · τ + ρ(c) g (71)where ρ(c) is generally a l<strong>in</strong>ear functionρ(c) = ρ 1 + c − c 1c 2 − c 1(ρ 2 − ρ 1 ) (72)It must be emphasized that this system does not satisfy the mass balance equation (1). Indeed,from equations (69) and (70), it is straightforward to show that∂ρdρ+ ∇ · (ρ v) = ∇ · [κ ∇˜µ] ≠ 0∂t dcHowever, if dρ/dc = cste, this equation can be written <strong>in</strong> a conservative form, which shows that(by <strong>in</strong>tegration over the entire volume), the total mass of the system is conserved (an importantproperty that many sharp <strong><strong>in</strong>terface</strong> numerical methods do not satisfy).Despite its simplicity, it can be shown that this system is ill-posed, <strong>in</strong> the sense that no monotonicallydecreas<strong>in</strong>g energy is associated to this system of equations; it thus violates the secondlaw of thermodynamics. This may be the price to pay to conserve the eas<strong>in</strong>ess of the numericalimplementation. . .However, it is possible to develop a diffuse <strong><strong>in</strong>terface</strong> model with a f<strong>in</strong>ite density contrast thatis thermodynamically coherent. One of the ma<strong>in</strong> issue is to def<strong>in</strong>e <strong>in</strong>compressibility flow whenthe velocity field is not solenoidal. Lowengrub and Trusk<strong>in</strong>ovsky [1998] showed that, <strong>in</strong> this case,it is relevant to use the thermodynamic <strong>in</strong>compressibility: the density is <strong>in</strong>dependent of the pressure.The ma<strong>in</strong> thermodynamic variables are the mass fraction c and the thermodynamic pressure Pand the relevant thermodynamic potential is not the free energy F but the specific Gibbs freeenthalpy g(c, P, ∇c) whose expression is:g(c, P, ∇c) = f 0 (c) + α 2 (∇c)2 +where the l<strong>in</strong>earity <strong>in</strong> P is due to the <strong>in</strong>compressibility assumption. The coefficient α is equivalentto a capillary coefficient but its dimension is not that of λ ([α] = [λ]/[ρ]).Lowengrub and Trusk<strong>in</strong>ovsky [1998] show that the correspond<strong>in</strong>g system of balance equationsis the follow<strong>in</strong>g:Pρ(c)whereρ dvdt∂ρ+ ∇ · (ρ v) = 0∂tρ dc = ∇ · (κ ∇˜µ)dt= −∇P − ∇ · (α ρ ∇c ⊗ ∇c) + ∇ · τ˜µ = df 0dc − P dρρ 2 dc − 1 ∇ · (ρ α ∇c)ρ27


This system is thermodynamically coherent and has been used to study p<strong>in</strong>ch-off and reconnectionof <strong><strong>in</strong>terface</strong>s for <strong>in</strong>stance [Lee et al., 2002a,b]. However, this sytem is much more coupledthan the <strong>in</strong>compressible model (or the thermodynamically <strong>in</strong>coherent model (69)-(71)). Indeed,the density ρ(c) appears <strong>in</strong> many terms and <strong>in</strong> particular <strong>in</strong> expression for the Korteweg stresstensor and <strong>in</strong> the Laplacian part of ˜µ. Moreover (and certa<strong>in</strong>ly most importantly), the pressureP appears <strong>in</strong> the expression for ˜µ, which <strong>in</strong>duces a complex coupl<strong>in</strong>g between the Cahn-Hilliardand momentum balance equation.4 <strong>Diffuse</strong> <strong><strong>in</strong>terface</strong> <strong>models</strong> and numerical simulation of mesoscopicproblems4.1 Numerical vs physical <strong><strong>in</strong>terface</strong> thicknessIn section 1.3, we showed that, <strong>in</strong> <strong>fluid</strong> <strong>mechanics</strong>, diffuse <strong><strong>in</strong>terface</strong> <strong>models</strong> are attractive fortwo ma<strong>in</strong> reasons: (i) they allow to study peculiar physical phenomena where sharp <strong><strong>in</strong>terface</strong><strong>models</strong> fail and (ii) they can be easily implemented numerically, which virtually elim<strong>in</strong>ates thedifficult and tedious numerical treatment of the mov<strong>in</strong>g boundaries. Whatever the reason, whenwe have to solve the equations of motion numerically, these equations are discretized <strong>in</strong> time andspace. In particular, if we consider the discretization <strong>in</strong> space, the <strong><strong>in</strong>terface</strong>s have to be capturedby the mesh. To illustrate this po<strong>in</strong>t, let us take a particular example of a regular mesh wherethe size of a mesh cell is ∆x <strong>in</strong> all directions. S<strong>in</strong>ce the cont<strong>in</strong>uous partial differential equationsare discretized on this mesh, all the spacial variations must be captured, so that the numericaltruncation errors do not make the system degenerate. In particular, the <strong><strong>in</strong>terface</strong> structure mustbe captured. This means that ∆x must be smaller than the <strong><strong>in</strong>terface</strong> thickness h. Typically, forstandard discretization schemes, one has ∆x ≃ h/4 so that the bi-Laplacian of the Cahn-Hilliardequation is well approximated. Even for higher order schemes, ∆x is of the order of h.The diffuse <strong><strong>in</strong>terface</strong> <strong>models</strong> presented <strong>in</strong> the previous sections (either for a liquid-vapor systemof for non-miscible phases) have been derived based on physical arguments: the ma<strong>in</strong> thermodynamicvariable are physical variables (mass density ρ or mass concentration c), the energyfunctional is physical and the equations of motion are based on physical first pr<strong>in</strong>ciples of conservation.Therefore, the <strong>in</strong>ternal structure of the <strong><strong>in</strong>terface</strong> is physical and <strong>in</strong> particular the <strong><strong>in</strong>terface</strong>thickness is also physical. This approach is relevant for any study where the <strong><strong>in</strong>terface</strong>s must bemodeled as cont<strong>in</strong>uous transition zones: coalescence and rupture of <strong><strong>in</strong>terface</strong>s, nucleation, etc. Inthese cases, the typical <strong><strong>in</strong>terface</strong> thickness is of the order of 10 −9 m and ∆x ≃ 10 −9 m. Therefore,assum<strong>in</strong>g that the maximum size of the mesh is of the order 10 6 mesh cells, the typical size of the3D doma<strong>in</strong> studied is about ( 10 −7) 3m 3 for a regular mesh. This is extremely small, even thoughit is relevant for this k<strong>in</strong>d of physical processes occur<strong>in</strong>g at the scale of an <strong><strong>in</strong>terface</strong>.However, when diffuse <strong><strong>in</strong>terface</strong> <strong>models</strong> are used for numerical convenience, if the physicalmodel is used as it is, one must restrict the studies to doma<strong>in</strong>s whose typical size is 10 −7 m. Now,<strong>in</strong> many applications, the typical size of the <strong>in</strong>clusions (bubbles or droplets) of the two-phase systemis much larger than this, by at least one order of magnitude. At this po<strong>in</strong>t, different strategiesare possible: (i) diffuse <strong><strong>in</strong>terface</strong> <strong>models</strong> are abandonned as well as their potential ease of use, (ii)adaptative mesh ref<strong>in</strong>ement techniques may be developed to capture the <strong>in</strong>terfacial zones or (iii)diffuse <strong><strong>in</strong>terface</strong> <strong>models</strong> are adapted to the numerical simulation of mesoscopic problems. In thelatter case, the adaptation of the model must be such that the value of the <strong><strong>in</strong>terface</strong> thickness isno longer dictated by physical arguments but rather by numerical arguments.The third solution will be studied <strong>in</strong> the subsequent sections. Nevertheless, the second solutiondeserves to be discussed. Indeed, we believe that it is a promis<strong>in</strong>g strategy because manyphysical phenoma occur close to the <strong><strong>in</strong>terface</strong> and a f<strong>in</strong>e discretization is therefore necessary tocapture them. However, us<strong>in</strong>g mesh ref<strong>in</strong>ement to capture the <strong>in</strong>ternal structure of an <strong><strong>in</strong>terface</strong>might not be the most relevant solution. Indeed, if one is <strong>in</strong>terested <strong>in</strong> problems whose typicalsize is that of a bubble or droplet (mesocopic scale), the sharp <strong><strong>in</strong>terface</strong> approximation is the most28


elevant. This means that all the physical processes of <strong>in</strong>terest occur at the scale of the <strong>in</strong>clusion(bubble or droplet) and not of the <strong><strong>in</strong>terface</strong> structure. Thus, there is a clear and justified scaleseparation. Now, us<strong>in</strong>g a mesh ref<strong>in</strong>ement technique for these problems means that a complexnumerical technique is used to capture phenomena of no physical <strong>in</strong>terest that have been <strong>in</strong>troducedto simplify the numerical implementation. Thus, seek<strong>in</strong>g for a way to use diffuse <strong><strong>in</strong>terface</strong><strong>models</strong> with an artificial <strong><strong>in</strong>terface</strong> thickness appears as the most relevant solution.4.2 Necessary modification of the diffuse <strong><strong>in</strong>terface</strong> <strong>models</strong>In the previous section we showed that, for many problems where the relevant scale is the radiusof an <strong>in</strong>clusion, the scale separation with the <strong><strong>in</strong>terface</strong> thickness is justified. In this case, it isalmost impossible to use physical diffuse <strong><strong>in</strong>terface</strong> <strong>models</strong> on a regular mesh: much too manymesh po<strong>in</strong>ts would be necessary only to capture the <strong>in</strong>terfacial zones. Therefore, the typical sizeof the <strong><strong>in</strong>terface</strong>s has to be adapted so that a reasonnable mesh can be used. In this case, the meshis more or less given, and therefore ∆x is given. Thus, the <strong><strong>in</strong>terface</strong> thickness h must be adaptedso that the <strong><strong>in</strong>terface</strong> structure can be captured by the mesh. The <strong><strong>in</strong>terface</strong> thickness h shouldtherefore be a free parameter whose value can be chosen arbitrarily. In the subsequent sections,we study if and how this is possible. We beg<strong>in</strong> with the van der Waals model and then we studythe Cahn-Hilliard model.4.3 Liquid-vapor flows with phase-change4.3.1 Modification of the parametersIn section 2.1.2, we showed that, with the van der Waals model, the <strong><strong>in</strong>terface</strong> thickness is givenby√1 λh =(73)ρ l − ρ v 2 Awhere λ is the capillary coefficient and A is a coefficient that characterizes the function W (ρ) andtherefore the free energy of the <strong>fluid</strong> F (ρ) (see equation (25)).The goal is that h can be chosen arbitrarily. Now, the <strong><strong>in</strong>terface</strong> thickness is a consequence ofthe diffuse <strong><strong>in</strong>terface</strong> model; it is not a primary parameter of the model but rather a secondaryparameter. The primary parameters of the model are ρ l , ρ v , λ and A. Therefore, these are theparameters on which one might have a degree of freedom to fix the value of h arbitrarily. Amongthe primary parameters, λ is the only “non-classical” parameter: all the others are <strong>in</strong>volved <strong>in</strong>the properties of the bulk phases. In particular, the parameter A is characteristic not only of thethermodynamic behavior of the <strong>fluid</strong> with<strong>in</strong> the <strong><strong>in</strong>terface</strong> but also with<strong>in</strong> the bulk phases: thefunction F 0 (ρ) (<strong>in</strong> which the parameter A appears) is valid for any value of ρ and <strong>in</strong> particular forthe values of ρ reached with<strong>in</strong> the bulk phases. In particular, it can be shown that the isothermalcompressibility of the bulk phases at saturation are given by( ) ∂P= 2 A ρ v (ρ l − ρ v ) 2 (74)∂ρv( ) ∂P= 2 A ρ l (ρ l − ρ v ) 2 (75)∂ρlThus, the only parameter clearly associated to the <strong><strong>in</strong>terface</strong> is λ and it appears as the mostobvious parameter that can be modified to <strong>in</strong>crease the <strong><strong>in</strong>terface</strong> thickness. Equation (73) showsthat λ should be <strong>in</strong>creased to <strong>in</strong>crease h.Now, we have also shown that the expression for the surface tension is the follow<strong>in</strong>g:σ = (ρ l − ρ v ) 36√2 A λ (76)29


P 0modifiedP eqρ vρ lρFigure 10: Modification of the equation of state P 0 (ρ) <strong>in</strong>duced by the <strong>in</strong>crease of the <strong><strong>in</strong>terface</strong>thickness.This expression shows that if λ is <strong>in</strong>creased while all the other parameters of the <strong>models</strong> are keptconstant, the value of σ is <strong>in</strong>creased by the same factor of h. Surface tension is an importantphysical parameter. Indeed, through the Laplace and Gibbs-Thompson relations (cf. equations(11) and (13)), this <strong>in</strong>terfacial property has an <strong>in</strong>fluence on bulk properties such as the pressure,the temperature and thus all the other bulk properties. Therefore, modify<strong>in</strong>g σ implies that theoverall flow at the mesoscopic scale (of <strong>in</strong>terest) is modified as well. This cannot be acceptable.To overcome this issue, at least one of the other parameters of the model has to be modifiedas well. The parameter A is chosen to be modified. Equations (31) and (33) show that, <strong>in</strong> order to<strong>in</strong>crease h and to keep σ constant, λ must be <strong>in</strong>creased and A must be decreased proportionally. It can beshown that, for given values of σ (physical) and h (numerical), λ and A are given byλ =A =32 (ρ l − ρ v ) 2 σ h12(ρ l − ρ v ) 4 σhThis analysis shows that the overall thermodynamic behavior of the <strong>fluid</strong> must be modifiedand not only its “non-classical” part <strong>in</strong> (∇ρ) 2 . This means <strong>in</strong> particular that the classical part ofthe equation of state F 0 (ρ) must be modified. The ma<strong>in</strong> consequences of this modification areanalyzed <strong>in</strong> the follow<strong>in</strong>g section.4.3.2 Consequences of this modificationIn the previous section, we showed that to make the <strong><strong>in</strong>terface</strong> thickness a free parameter of thesystem, it is necessary to modify the thermodynamic behavior of the <strong>fluid</strong>. In particular, it isnecessary to modify the “classical” part of the free energy F 0 (ρ). This is an unexpected consequenceof the procedure. Indeed, the <strong>in</strong>itial goal was to modify only the <strong><strong>in</strong>terface</strong> thickness.By do<strong>in</strong>g this, we implicitly accepted to violate the detailed physics of the <strong><strong>in</strong>terface</strong> layer. Butwe wanted to modify only <strong><strong>in</strong>terface</strong> structure without modify<strong>in</strong>g the outer bulk properties. Theabove analysis shows that it is not possible. Now, the issue is to study whether the modificationof the free energy F 0 (ρ) has important consequences on the behavior of the two-phase system atthe mesoscopic scale.This study is rather difficult. Very often, this is done through an asymptotic analysis of thesystem. The goal is to determ<strong>in</strong>e to which sharp <strong><strong>in</strong>terface</strong> model the diffuse <strong><strong>in</strong>terface</strong> model isequivalent. In the particular case where a diffuse <strong><strong>in</strong>terface</strong> model is used ma<strong>in</strong>ly for numericalconvenience, the sharp <strong><strong>in</strong>terface</strong> model is known (cf. section 1.2) and we want the diffuse <strong><strong>in</strong>terface</strong>model to mimic this given sharp <strong><strong>in</strong>terface</strong> model as well as possible. In particular, we aim atrecover<strong>in</strong>g the same boundary conditions made of <strong>in</strong>terfacial mass, momentum, energy balanceequations and, at first approximation, local thermodynamic equilibrium of the <strong><strong>in</strong>terface</strong>. The issueis thus to determ<strong>in</strong>e whether the modification of the parameter A modifies these boundaryconditions.30


The modification of A does not modify any property at saturation. In particular, the saturationpressure and the densities at saturation are not modified. However, it modifies the variations ofthe pressure with the density around the states at saturation ρ v and ρ l . In particular, it modifiesthe compressibility of the bulk phases. Compressibility has an effect on the motion of the <strong>fluid</strong>only if the typical velocity of the <strong>fluid</strong> is of the order of magnitude of the speed of sound (thatdepends on the compressibility of the <strong>fluid</strong>). In many applications, the bulk velocities are verysmall compared to the speed of sound and, even though the latter is decreased by decreas<strong>in</strong>g A,this rema<strong>in</strong>s valid.However, thermal expansion effects are important to account for because they are at the orig<strong>in</strong>of natural convection. The expression for the thermal expansion coefficient isκ T = 1 ρ(∂P/∂ρ) T(∂P/∂T ) ρThis expression shows that, if (∂P/∂ρ) T is decreased, (∂P/∂T ) ρ must be decreased as well to conservethe value of κ T . Now, the decrease of (∂P/∂T ) ρ is equivalent to the decrease of (dP sat /dT ).Because of the Clapeyron relationdP satdT = LT (v v − v l )<strong>in</strong> order the keep the value of L constant, it is necessary to modify the value of T . Actually, athorough analysis of this issue [Fouillet, 2003] shows that only the value of the temperature ofreference has to be modified. S<strong>in</strong>ce only temperature differences are important, the <strong>in</strong>crease ofthe reference temperature is not very important. Moreover, it has been shown that this modificationma<strong>in</strong>ly <strong>in</strong>fluences the time scale at which the <strong>in</strong>terfacial zone gets back to the saturationtemperature. This time is physically extremely small and an <strong>in</strong>crease of the time generally barelyhas any <strong>in</strong>fluence.This short analysis shows that the issue of the numerical <strong>in</strong>crease of the <strong><strong>in</strong>terface</strong> thickness isnot trivial. It also shows that, even though the <strong>in</strong>itial goal was to modify only a property of the<strong>in</strong>ternal structure of the <strong><strong>in</strong>terface</strong>, it lead to modify important properties of the bulk phases aswell; it actually lead to modify all the variation of the equation of state P (ρ, T ). This is ma<strong>in</strong>lybecause the graph of this function (or equivalently the function F 0 (ρ, T )) <strong>in</strong>fluences bulk propertiesas well as the <strong>in</strong>ternal structure of the <strong><strong>in</strong>terface</strong>.However, us<strong>in</strong>g these modifications, the van der Waals model can be used to simulate complexproblems such as nucleate boil<strong>in</strong>g on a heated surface [Jamet and Fouillet, 2005] as shown<strong>in</strong> figure 11.This model can be extended to dilute b<strong>in</strong>ary mixtures as shown <strong>in</strong> figure 12. The <strong>in</strong>fluenceof the addition of a small amount of an extra component has been shown to have an important<strong>in</strong>fluence on the wall heat exchange coefficient (cf. figure 13), which is commonly observed experimentally.Nevertheless, it is very difficult to use this model to get quantitative results. One of thema<strong>in</strong> issues is gravity. Indeed, it has been shown that all the pressure scales had to be modified:(∂P/∂ρ) and (∂P/∂T ). These modifications are acceptable as long as no external pressurescale is imposed. But gravity imposes an external scale of pressure variation. To illustrate theissue, let us consider a vapor bubble on a heated wall. Because of gravity, the pressure at the topof the bubble is lower than the pressure at the bottom of the bubble. The local thermodynamicequilibrium of the <strong><strong>in</strong>terface</strong> is determ<strong>in</strong>ed by the Clapeyron relation which determ<strong>in</strong>es the equilibrepressure as a function of the local temperature. So that the <strong><strong>in</strong>terface</strong> rema<strong>in</strong>s at equilibrium,the temperature must vary <strong>in</strong> the direction of the gravity field: lower at the top of the bubble thanat its bottom:ρ g∇T =(dP sat /dT ) ≃ ρ g(∂P/∂T ) ρ31


Figure 11: Numerical simulation of nucleate boil<strong>in</strong>g us<strong>in</strong>g the van der Waals’ diffuse <strong><strong>in</strong>terface</strong>model [Jamet and Fouillet, 2005]. The color field represents the temperature and the <strong><strong>in</strong>terface</strong>corresponds to iso-contours of the density field.Figure 12: Numerical simulation of nucleate boil<strong>in</strong>g us<strong>in</strong>g an extension of the van der Waals’diffuse <strong><strong>in</strong>terface</strong> model to b<strong>in</strong>ary mixtures [Jamet and Fouillet, 2005].The color field represents themass fraction of a dilute substance and the <strong><strong>in</strong>terface</strong> corresponds to iso-contours of the densityfield.S<strong>in</strong>ce (∂P/∂T ) ρ is decreased because of the <strong>in</strong>crease of the <strong><strong>in</strong>terface</strong> thickness, ∇T is <strong>in</strong>creased.If the thermal conditions of the system more or less impose the temperature gradient (if the heatflux is imposed at the heated wall for <strong>in</strong>stance), the disequilibrium of the bubble <strong><strong>in</strong>terface</strong> ismodified and the dynamics of the bubble as well.32


DT (K)12108642mixturepure <strong>fluid</strong>h (J/m^2/s/K)16000014000012000010000080000600004000020000mixturepure <strong>fluid</strong>00 0.5 1 1.5 2 2.5 3t (s)00 0.5 1 1.5 2 2.5 3t (s)Figure 13: Comparison of the mean temperature wall of of the <strong>in</strong>stantaneous mean heat flux withand without the presence of a dilute substance [Jamet and Fouillet, 2005].4.4 Two-phase flows of non-miscible <strong>fluid</strong>sIn the case of non-miscible phases, the issue is the same: how to make the <strong><strong>in</strong>terface</strong> a free parameterwithout modify<strong>in</strong>g the flow at the mesoscopic scale? The solution is the same as <strong>in</strong> theliquid-vapor case: <strong>in</strong>crease λ and decrease A proportionally. However, <strong>in</strong> this case, the consequencesare different. The variations of the classical chemical potential µ 0 (c) must be modified.We showed <strong>in</strong> the previous section that this modification might not be critical as long as noexternal scale of variation of the chemical potential is imposed. In common two-phase flow applications,it is rare that any external scale of variation of the chemical potential is imposed andthe situation is thus easier that <strong>in</strong> the liquid-vapor case. However, there does exist a difficulty. Indeed,the mobility κ is related to the time scale at which a system goes back to equilibrium. S<strong>in</strong>cethe <strong><strong>in</strong>terface</strong> thickness is modified, if κ is not modified, the time scale at which the <strong>in</strong>terfacial zonegoes back to equilibrium is modified. Therefore, κ must be <strong>in</strong>creased to ensure that the <strong><strong>in</strong>terface</strong>keeps close to equilibrium. F<strong>in</strong>d<strong>in</strong>g the optimal value for κ is not straightforward. Now, <strong>in</strong> section3.1.2 we showed that the chemical potential of equilibrium of a spherical <strong>in</strong>clusion dependson the radius of curvature of the system. Therefore, if two spherical <strong>in</strong>clusion of different radiiat put close to each other, because of this difference of chemical potentials, a diffusion mass fluxdevelops <strong>in</strong> between the <strong>in</strong>clusions, mak<strong>in</strong>g the smaller <strong>in</strong>clusion shr<strong>in</strong>k and the bigger expand(cf. figure 14). This diffusion mass flux is proportional to the difference of curvature of the <strong>in</strong>clusionsand to the mobility κ, the latter be<strong>in</strong>g modified. Thus the time scale at which the processoccurs is decreased. This problem is generally overcome by mak<strong>in</strong>g κ vary so that its value <strong>in</strong>the bulk phases vanishes, thus elim<strong>in</strong>at<strong>in</strong>g this “parasitic coarsen<strong>in</strong>g”. However, it can be shownthat, because of numerical truncation errors (and also other fundamental reasons that cannot bedeveloped here) this “parasitic coarsen<strong>in</strong>g” cannot be totally elim<strong>in</strong>ated.Figure 14: Mass diffusion between <strong>in</strong>clusions of different sizes. The color field represents thegeneralized chemical potential and the <strong><strong>in</strong>terface</strong>s are represents by iso-contours of the mass fraction.Even though the issue of the numerical <strong>in</strong>crease of the <strong><strong>in</strong>terface</strong> thickness might be less critical33


<strong>in</strong> the Cahn-Hilliard model than <strong>in</strong> the van der Waals model, it is nevertheless a real difficulty toensure that these modifications do not <strong>in</strong>volve modifications on the mesoscopic characteristics ofthe flow.This analysis shows that the adaptation of physical diffuse <strong><strong>in</strong>terface</strong> <strong>models</strong> to mesoscopicproblems is rather difficult and tricky. We showed <strong>in</strong> particular that it is important to know thesharp <strong><strong>in</strong>terface</strong> model that the diffuse <strong><strong>in</strong>terface</strong> model must mimic. It is <strong>in</strong>deed a reference thathelps to develop the “equivalent” diffuse <strong><strong>in</strong>terface</strong> model. Moreover, we showed that the difficultycomes from the fact that we use only physical variables (mass density ρ or mass fractionc) and that these physical variables are important to characterize (i) the <strong><strong>in</strong>terface</strong> structure (thatis aimed at be<strong>in</strong>g modified) and (ii) bulk phases properties (pressure, chemical potential, etc).We showed that modify<strong>in</strong>g one feature without modify<strong>in</strong>g the other is difficult and tricky. Thesystem lacks degrees of freedom. This degree of freedom can come from the <strong>in</strong>troduction of anotherparameter whose ma<strong>in</strong> goal would be to characterize only the <strong><strong>in</strong>terface</strong> structure and notthe bulk properties. The phase-field variable ϕ often used <strong>in</strong> other applications of diffuse <strong><strong>in</strong>terface</strong><strong>models</strong> can be <strong>in</strong>terpreted as such a variable. It has recently been shown [Jamet and Ruyer,2004] that such a phase-field model<strong>in</strong>g is possible for liquid-vapor flows. The <strong>in</strong>troduction of thephase-field <strong>in</strong>deed allows to easily decouple the <strong><strong>in</strong>terface</strong> properties from the bulk properties.ReferencesJ. U. Brackbill, D. B. Kothe, and C. Zemach. A cont<strong>in</strong>uum method for model<strong>in</strong>g surface tension.J. Comp. Phys., 100:335–354, 1992.J. W. Cahn and J. E. Hilliard. Free energy of a nonuniform system. I. <strong>in</strong>terfacial free energy. J.Chem. Physics, 28(2):258–267, 1958.J. W. Cahn and J. E. Hilliard. Free energy of a nonuniform system. II. thermodynamic basis. J.Chem. Physics, 30(5):1121–1124, 1959a.J. W. Cahn and J. E. Hilliard. Free energy of a nonuniform system. III. nucleation <strong>in</strong> a twocomponent<strong>in</strong>compressible <strong>fluid</strong>. J. Chem. Physics, 31(3):688–699, 1959b.J.-M. Delhaye, M. Giot, and M. L. Riethmuller. Thermohydraulics of two-phase systems for <strong>in</strong>dustrialdesign and nuclear eng<strong>in</strong>eer<strong>in</strong>g. Hemisphere Publish<strong>in</strong>g Corporation, 1981.F. Dell’Isola, H. Gou<strong>in</strong>, and G. Rotoli. Nucleation of spherical shell-like <strong><strong>in</strong>terface</strong>s by secondgradient theory: Numerical simulations. Eur. J. Mech. B/Fluids, 15(4):545–568, 1996.J. E. Dunn and J. Serr<strong>in</strong>. On the thermodynamics of <strong>in</strong>tersticial work<strong>in</strong>g. Arch. Rational Mech.Anal., 88:88–133, 1965.C. Fouillet. Généralisation à des mélanges b<strong>in</strong>aires de la méthode du second gradient et application à lasimulation numérique directe de l’ébullition nucléée. Thèse de doctorat, Université Paris VI, 2003.C. Fouillet, D. Jamet, and D. Lhuillier. A cont<strong>in</strong>uous <strong><strong>in</strong>terface</strong> model for the direct numerical simulationof phase-change <strong>in</strong> two-component liquid-vapor flows. In 2002 Jo<strong>in</strong>t ASME/EuropeanFluid Eng<strong>in</strong>eer<strong>in</strong>g Division Summer Conference, Montreal, Canada, July 14-18, 2002.G. P. Galdi, D. D. Joseph, L. Preziosi, and S. Rionero. Mathematical problems for miscible andcompressible <strong>fluid</strong>s with korteweg stresses. European Journal of Mechanics B Fluids, 10(3):253–267, 1991.H. Gou<strong>in</strong>. Energy of <strong>in</strong>teraction between solid surfaces and liquids. J. Phys. Chem. B, 102:1212–1218, 1998. doi: 10.1021/jp9723426.34


D. Jacqm<strong>in</strong>. Calculation of two-phase Navier-Stokes flows us<strong>in</strong>g phase-field model<strong>in</strong>g. J. Comp.Phys., 155:1–32, 1999.D. Jacqm<strong>in</strong>. Contact-l<strong>in</strong>e dynamics of a diffuse <strong>fluid</strong> <strong><strong>in</strong>terface</strong>. J. Fluid Mech., 402:57–88, 2000.D. Jamet, J. U. Brackbill, and D. Torres. On the theory and computation of surface tension: Theelim<strong>in</strong>ation of parasitic currents through energy conservation <strong>in</strong> the second gradient method.J. Comp. Phys., 182:262–276, 2002. doi: 10.1006/jcph.2002.7165.D. Jamet and C. Fouillet. Direct numerical simulations of nucleate boil<strong>in</strong>g flows of b<strong>in</strong>ary mixtures.In 11th International Topical Meet<strong>in</strong>g on Nuclear Thermal-Hydraulics (NURETH-11), Popes’Palace Conference Center, Avignon, France, October 2-6, 2005. Paper # 364.D. Jamet and P. Ruyer. A quasi-<strong>in</strong>compressible model dedicated to the direct numerical simulationof liquid-vapor flows with phase-change. In 5th International Conference on MultiphaseFlows, ICMF’04, Yokohama, Japan, May 30-June 4, 2004.D. J. Korteweg. Sur la forme que prennent les équations du mouvement des <strong>fluid</strong>es si l’on tientcompte des forces capillaires causées par des variations de densité considérables mais cont<strong>in</strong>ueset sur la théorie de la capillarité dans l’hypothèse d’une variation cont<strong>in</strong>ue de la densité.Arch. Néerl. Sci. Exactes Nat., 6:1–24, 1901.H. G. Lee, J. S. Lowengrub, and J. Goodman. Model<strong>in</strong>g p<strong>in</strong>choff and reconnection <strong>in</strong> a hele-shawcell I: The <strong>models</strong> and their calibration. Phys. Fluids, 14(2):492–513, 2002a.H. G. Lee, J. S. Lowengrub, and J. Goodman. Model<strong>in</strong>g p<strong>in</strong>choff and reconnection <strong>in</strong> a hele-shawcell II: Analysis and simulation <strong>in</strong> the nonl<strong>in</strong>ear regime. Phys. Fluids, 14(2):514–545, 2002b.J. Lowengrub and L. Trusk<strong>in</strong>ovsky. Quasi-<strong>in</strong>compressible cahn-hilliard <strong>fluid</strong>s and topologicaltransitions. Proc. R. Soc. Lond. A, 454:2617–2654, 1998.Y. Rocard. Thermodynamique. Masson, Paris, 1967.J. S. Rowl<strong>in</strong>son and B. Widom. Molecular theory of capillarity. Oxford University Press, New York,1982.P. Seppecher. Mov<strong>in</strong>g contact l<strong>in</strong>e <strong>in</strong> the Cahn-Hilliard theory. Int. J. Engng. Sci., 34(9):977–992,1996.L. Trusk<strong>in</strong>ovsky. K<strong>in</strong>ks versus shocks. In R. Fosdick J. E. Dunn and M. Slemrod, editors, ShockInduced Transitions and Phase Structure <strong>in</strong> General Media, IMA Series <strong>in</strong> Math. and its Appl., pages185–229. Spr<strong>in</strong>ger Verlag, 1993.van der Waals. Thermodynamische Theorie der Kapillarität unter Voraussetzung stetiger Dichteanderung.Z. Phys. Chem., 13:657–725, 1894. English translation <strong>in</strong> J. Statist. Phys., 20, 197.35

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

Saved successfully!

Ooh no, something went wrong!