Interface Tracking Methods (ITM) are used for free-surface and interfacial flows which involve two or
more immiscible fluids (e.g., liquid-liquid or liquid-gas systems) separated by sharp interfaces which
evolve in time. Interface tracking/capturing schemes are methods that are able to locate
the interface by capturing the interface by keeping track of the evolution of an appropriate
field.
The formulation is based on solving a single set of transport equations for the whole
computational domain and treating the two phases as a single fluid with variable material properties.
Changes in these properties are taken into account by advecting the phase-indicator function
χ(x,t):
(1.1)
where the interface velocity uI reduces in isothermal flow conditions (no phase change) to the local fluid
velocity u.
Mass exchange terms and capillary forces are represented as momentum and energy sources,
described as delta functions or smoothed gradients at or across the interface. The mass,
momentum and energy balance equations for incompressible interfacial flow are written
as
where γ is the surface tension coefficient, κ is the interface curvature, n is the normal vector to the
interface, δI is a smoothed Dirac delta function centered at the interface, T is the temperature, Cp is
the heat capacity of the fluids, λ is the heat conductivity, and is the volumetric heat source. Material
properties are updated using χ:
1.2 The VOF methodology
The VOF approach relies on the definition of the liquid volume-fraction field occupied by
one of the phases within volume V , and is conventionally denoted in a discrete form by
Fijk:
(1.9)
In the VOF context, Eq. (1.1), therefore, represents the evolution of the liquid volume-fraction,
identifying flow regions containing pure liquid (where Fijk = 1) from pure gas flow regions (where
Fijk = 0). Interfacial cells are such that 0 < Fijk< 1. The VOF method does not amount solely to the
solution of:
(1.10)
It requires accurate algorithms for advecting the volume fraction function so as to preserve the thickness
of the interface. In order to keep the interface sharp a special advection scheme referred to as the
Compressive Interface Capturing Scheme for Arbitrary Meshes (CICSAM) (Ubbink, 1997; Ubbink &
Issa, 1999) is used.
1.3 The Level Set method
1.3.1 The scheme
The Level Set approach (Osher & Sethian, 1988) consists in solving a hyperbolic equation:
(1.11)
to track the interface on a fixed Eulerian grid, where ϕ(x,t) is a smooth signed-distance function
referring to the shortest distance to the front. Negative values correspond to one of the fluids and
positive values to the other. The exact location of the interface corresponds to the zero level of ϕ. To
update material properties like density, viscosity and thermal conductivity, a Heaviside function H(ϕ) is
introduced and defined by
(1.12)
This implies that ϕijk is linked in discrete form to the liquid volume fraction field Fijk
through
(1.13)
A modified Heaviside function denoted by H(ϕ) is employed to smooth the physical properties across
an interface thickness of 2δ.
(1.14)
where Hk(ϕ) is the modified Heaviside function defined by
(1.15)
The same modified Heaviside function is concurrently employed to determine the surface
tension.
1.3.2 Re-distancing scheme
As numerical errors cause the contours of the Level Set field to deform as the phase moves, a
re-distancing algorithm is required to regularise the function: advecting the initial distance function
ϕ(x,0) will not be maintained as such. An extra re-distancing algorithm preserving |∇ϕ| = 1 around the
zero level of ϕ is required.
The scheme of Sussman et al. (1994) consists of advecting an intermediate equation for a correction
field d(x,t) during a period of intermediate time :
(1.16)
with sgn(ϕ) = 2H(x) - 1 denotes the signum function, d0(x,t) = ϕ(x,t), and
(1.17)
The last term in the above expression is added to keep the bubble volume constant during the
reinitialisation procedure. The equation is iterated until |∇d| = 1. The corrected ϕ–field is then obtained
by setting ϕ(x,t) = d(x,ε), where ε is the time elapsed for Eq. (1.16) to reach convergence. In TransAT,
Eq. (1.16) is solved after each advection step of the Level Set equation using the non-oscillatory
(WENO) 3rd order scheme.
1.3.3 Fast-marching re-distancing
This is an extreme case of the narrow band method, in that the band in question is taken as a single
cell. Of interest in Level Set methods is to preserve the normal distribution of the iso-contours of
function ϕ(x) through the Eikonal equation:
(1.18)
with boundary conditions ϕ(x) = g(x),x ∈ Γ ⊂ Rn. In the simplest case where the speed function f(x)
is a constant, usually equal to one, then the solution of the Eikonal equation above will represent the
shortest distance from a point x to the zero distance curve, usually given by Γ, that is if g(x) = 0. To
construct an approximation to the Eikonal equation for a grid point x using the surrounding points xi,
xn, first define the unit directional vector pointing from point xi to point x as Pi = (x-xi)∕||x-xi||.
Then let vi(x) be the value of the approximate directional derivative in direction Pi and point
x.
We then solve a quadratic equation for the distance ϕ(x) in the form
(1.19)
where Q is defined as Q = (PPT)-1 and the vectors a and b by their respective constituents ai and bi.
This equation is then solved to find two roots of which the largest one is taken as a possible new
distance for ϕ(x). This new distance can only be accepted if the update comes from within the polygon
spanned by x,xi,c…,xn and if the calculated distance value is smaller than the old one. Thus we only
update if ϕ(x) < ϕold(x).
1.3.4 Thin-film boundary conditions
For flows with very low Capillary numbers, there exists a thin water film between the gas phase and the
wall. A thin-film boundary condition can then be defined. The viscosity at the film–boundary wall is set
so as to satisfy the shear according to Couette flow. For a liquid film condition we get (see Fig. 1.1),
Figure 1.1: Sketch for thin-film boundary condition
(1.20)
The interface velocity (tangential to the wall) uI is then given as,
(1.21)
The effective viscosity is then set according to the following equation,
Marangoni force can be modelled, which will make the surface tension coefficient depend on
temperature
(1.24)
where, σ0 is the surface tension coefficient at temperature T0, and Tcrit is the temperature of the
material at which surface tension is zero.
The thermo-capillary (Marangoni) contribution to the Navier-Stokes equation is the projection
tangential to the interface of the shear stress generated by the temperature dependence of the surface
tension and writes:
(1.25)
1.3.6 Dynamic contact angle treatment
The treatment of wetting dynamics is based on the physical forces associated with triple lines (Friess &
Lakehal, 2004; Narayanan & Lakehal, 2006). An additional triple-line contact force (ci) is included in
the momentum equation; this extended momentum equation provides a physically adequate description
of wetting dynamics, eliminating the need for any particular boundary condition specifying the contact
angle. The triple-line force (ci) is based on a consideration of interfacial free energy. Accordingly, it
contains only two parameters: the interfacial tension between the fluids γ and the equilibrium contact
angle θeq.
(1.26)
where θdy is the instantaneous dynamic contact angle, δt is a smooth Dirac delta function vanishing
everywhere except near the triple lines, with the property that for any volume V , the integral ∫VδtdV
is equal to the length of the triple line segment contained in V , and b is the unit vector normal to the
triple line and parallel to the wall surface. The triple-line force, is obtained by considerations
similar to the derivation of Youngs Law and is referred to as the unbalanced Young force
(De Gennes, 1985).
In addition, the shear stress is assumed to have a logarithmic profile as it nears the contact line
(Qian et al., 2004). At distances less than a slip length from the triple line, full slip is assumed. In the
finite volume containing the triple line, the integrated shear stress is applied. The slip length is taken to
be a small value ≈ 10-9. This formulation resolves the stress singularity that arises due to the no–slip
boundary condition.
1.4 Phase change modelling
In the presence of phase change, equation (1.11) is rewritten
(1.27)
where ṁ′′ stands for the mass transfer per unit area per unit time. A new source term also appear in
the velocity divergence equation (1.2), which becomes
(1.28)
Finally, a new source term is added to the energy equation (1.4)
(1.29)
It now remains to model the mass transfer term ṁ′′.
1.4.1 Direct phase change
When using direct phase change, mass transfer is given by the formula
(1.30)
(1.31)
with the latent heat.
1.4.2 Correction of saturation temperature
The saturation temperature can be locally adjusted to take into account the effect of curvature, using
the equation
(1.32)
where, F(ρ) = ρG-1 + ρL-1, κ is the local curvature, σ is the local surface tension coefficient, and L is
the latent heat of phase change.
1.4.3 Correlation-based models
Different models are based on global correlation. In this case, local ṁ do not vary in space and time, but
the global value will change because of the changes of interfacial area.They are all based on the same
expression
(1.33)
where NuL is the liquid Nusselt number, is the latent heat and Tsat is the saturation temperature.
Tsc and LL,ṁ are given parameters. NuL can be given by different correlations.
is the liquid Reynolds number, and uL,ṁ is a correlation velocity.
Other correlations have also been proposed by Lim et al. (Lim et al., 1984):
(1.36)
for smooth interface, and
(1.37)
for wavy interface. In these expressions,
(1.38)
is the liquid Reynolds number.
1.4.4 Interface Mass Transfer (IMT)
This section describes the interfacial mass transfer models that have been implemented in TransAT
(Lakehal & Labois, 2011a). For the sake of simplicity, all of them will be written in the
form
(1.39)
where ṁ is the mass transfer rate, h is the heat transfer coefficient, u′ is a characteristic velocity of the
flow, Pr is the Prandtl number, Sc is the Schmidt number and Ret a turbulent Reynolds
number.
where C = 0.35 if Sc (or Pr) < 4, 0.45 otherwise. This model is adapted for low Reynolds number.
It is also implemented with modifications, for the small eddy context
(1.41)
This model is adapted for very high turbulent Reynolds numbers. Finally, an adaptive surface
divergence version has been implemented
(1.42)
with
(1.43)
where Ret,int is the turbulent Reynolds number at the interface. This model is adapted for high
Reynolds numbers.
The last model from (Coste, 2004) is inspired by the renewal model and reads
(1.48)
A turbulent Reynolds number is needed for the interfacial mass transfer model. Several choices are
available in TransAT:
The usual Reynolds number can be written Ret = . It is taken in the gas core
flow. A length scale must be associated to this Reynolds number, it is computed as
the weighted-average (in the gas side) of ε, with the characteristic velocity being
u∞ = min(|u|,Cμ1∕4k1∕2)
As an alternative to Ret, another Reynolds number based on turbulence is available,
Ry = . It is also evaluated in the gas core flow.
The shear Reynolds number reads Re⋆ = with u⋆ = the shear velocity at the
interface, ϕ the distance of the point to the interface. u⋆ and ϕ are taken in the gas near
the interface, where u⋆ is at its maximum. u⋆ is also taken as the associated characteristic
velocity.
Because they need the variables k and ε to be evaluated, ReT and Ry can be used only when the
k -ε turbulence model is enabled. Otherwise, Re⋆ together with the surface renewal model will be used.
The following table gives the available interfacial mass transfer models available for the different
Reynolds numbers.
turbulence model
Non-dimensional number
available IMT model
k - ε
“ReT”
”surfacedivg”,”SD_small_eddy,”SD_adaptive”,
”Hughes-Duffey”,”Coste”,”small_eddy”
k - ε
“Ry”
”surfacedivg”,
”Hughes-Duffey”,”Coste”,”small_eddy”,
k - ε, LES,
“Re⋆”
”surface_renewal”
laminar flow
Table 1.1: Interfacial mass transfer models available in TransAT.
We consider a mixture of two incompressible phases. Thus, the following equations are solved (Drew &
Passman, 1998):
(2.1)
(2.2)
where the subscript G denotes the gas phase, m denotes a mixture property, and T stands for mixture
turbulent property. In the following, the subscript L will also be introduced for the liquid phase. αG is
the gas volume fraction and ρm = αGρG + αLρL is the density of the mixture. Vm is the velocity of the
mixture, pm the mixture pressure, and Tm is the mixture temperature. SαG and Smomdrift stand for the
source terms due to mass transfer and drift velocities in the mass and momentum equations,
respectively.
The phasic ensemble averaged velocities split into mixture and drift velocities as follows,
and the phasic drift velocities are related to each other by,
(2.6)
where Yk = αkρk∕ρm is the mass fraction of the phase k. The slip velocity defined as Vsi = VGi-VLi
can be further related to the gas drift velocity as,
(2.7)
2.1.1 Homogeneous Model
In the homogeneous model, it is considered that both phases have the same velocity (Vkd = 0). The
equations can then be simplified to,
(2.8)
Only the mixture equations are solved. The divergence of the mixture velocity is zero.
(2.9)
2.1.2 Algebraic Slip Model
In this case, an algebraic model is used to estimate the relative velocity between the two
phases or between the mixture and the gas phase. Thus, several drift velocity models have
been implemented. With this notation, the algebraic slip model (ASM) can be written as
follows,
(2.10)
(2.11)
The divergence of the mixture velocity is given by,
(2.13)
Models for gas drift velocity V Gd must now be derived.
2.1.3 Algebraic slip model with drag forces
Ishii & Mishima (1984) have written the drag-induced momentum transfer term as
(2.14)
where CD is the drag coefficient and Rb the bubble radius.
Then, Manninen & Taivassalo (1996) derived from the conservation equations the following
expression, neglecting diffusion terms due to turbulent fluctuations and assuming the local
equilibrium.
(2.15)
From these equations, an expression for the gas drift velocity can be found, assuming that all
bubbles have the same radius Rb.
(2.16)
where Reb = is the bubble Reynolds number. Introducing the friction factor , the drag
coefficient can be written
(2.17)
and the drift velocity is rewritten
(2.18)
Friction factor remains to be modelled. In the Stokes Regime, it is simply
(2.19)
This model is adapted for very small bubbles Reynolds number, usually less than 2.
Clift (cited in Crowe et al. (1998)) developed a model that is adapted to bubble Reynolds numbers
up to 1000. In this model, friction factor is written
(2.21)
The model derived by Tomiyama (1998a) also takes into account the shape of the bubbles, using the
Eötvös number Eo = , where ||g|| is the gravitational acceleration and
σ the surface tension. The friction factor is then given (for highly contaminated system
here):
(2.22)
2.1.4 Algebraic slip model with drag and lift forces
Momentum transfer term 2.14 can then be rewritten as
(2.24)
and solving this equation will give the gas drift velocity. In this expression, lift coefficient is still
unknown. It can be taken as a constant, with values of (Drew & Passman, 1998) or (Drew &
Lahey, 1987) proposed in the literature for bubbly flows. Tomiyama (1998b) derived a more complex
model, in accordance with deformable bubbles correlation. It evaluates the long axis dH of a deformable
bubble, and proposes to compute the lift coefficient based on this axis. The model can then be
written
(2.28)
2.1.5 Wall lubrication force
When the lift force is used an unrealistic accumulation of small bubbles at the wall could appear. To
counterbalance this effect an additional wall lubrication force is typically included. The expression of the
wall lubrication force is given by Rzehak et al. (2012):
(2.29)
where nW is the unit vector perpendicular to the wall, pointing into the fluid. The second part of the
force was added to keep in account the fact that the lubrication force acts only close to the wall in a
exponentially way and not at the center of the vessel. Momentum transfer term 2.14 can then be
rewritten as:
(2.30)
In TransAT, different models are available. According to Antal et al. (1991), the expression of the
CWL coefficient was derived based on calculation of the flow field around a two dimensional circular
bubble in potential flow as:
(2.31)
Based on the observation of single bubble trajectories, Tomiyama et al. (1995) proposed this
formulation:
(2.32)
with the coefficient CW that depends on the Eotvos number. According to the Tomiyama
formulation, the coefficient CW of the Eq. 2.32 is:
(2.33)
Another correlation of the CW, depending on the Eotvos number, was proposed by Hosokawa
et al. (2002) as:
(2.34)
Frank et al. (2004) generalized the Tomiyama model to produce the Frank Wall Lubrication force
model, that is given by:
(2.35)
where CW is the coefficient defined in 2.33, that preserves the same dependence on Eotvos number as
the Tomiyama model.
2.1.6 Turbulent dispersion
The drift velocity models of the previous sections do not take into account the effect of the turbulence.
In this aim, Ishii (1975) proposed a term to be added to the gas drift velocity:
(2.36)
where νT is the turbulent kinematic viscosity and PrT is the turbulent Schmidt number.
2.2 Phase change modelling
Phase change phenomenon is divided into phase change happening in the bulk and phase change
happening near a wall (heated or cooled). The latter is called heterogeneous phase change.
2.2.1 Phase change in the bulk
Phase change happening in the bulk of the fluid (away from the flow boundaries) can be
phenomenologically split into different phase change modes, such as,
Dispersed phase change
Interfacial phase change
Cavitation
Homogeneous condensation
where the first two are thermally limited phase change phenomena and the last two are inertial phase
change phenomena where the spontaneous phase change is related to rapid pressure changes. Currently,
homogeneous condensation models are not available in TransAT.
Dispersed-phase phase change
This mode of phase change is defined as phase change that happens under thermal non-equilibrium
under conditions where the phases are well mixed. The phase change rate is limited by the temperature
difference (T - Tsat and the interfacial area density. The model available in TransAT is based on the
Ranz-Marshall Nusselt number model (Ranz & Marshall, 1952). The mass transfer rate (kg∕m3) is
given as,
where ṁ′′′ is the volumetric phase change rate in kg∕(m3s), q′′ is the interfacial heat flux, a is the
interfacial area density, L is the latent heat of vaporization or condensation, Nu is Nusselt number, dp is
the droplet/bubble diameter, Re is the droplet Reynolds number based on the slip velocity and Tsat is
the saturation temperature at the local pressure. The bubble/droplet diameter is specified as an input.
The system can be forced into phase/thermal equilibrium by specifying ever lower values for
dp.
Interfacial phase change
The main difference between the interfacial and the dispersed phase change models is the fact that for
conditions where the phases are separated, one can obtain an estimate for the real interfacial area using
the gradient of the volume fraction. Interfacial phase change has been split into two cases in
TransAT,
1.
Interfacial thermal diffusion phase change model, and
2.
Interfacial shear phase change model.
The interfacial shear phase change model is based on the proposal by Banerjee (Lakehal &
Labois, 2011b) where,
where ṁ′′′ is the volumetric phase change rate in kg∕(m3s), q′′ is the interfacial heat flux, a is the
interfacial area density, L is the latent heat of vaporization or condensation, Nu is Nusselt number, Reτ
is the interfacial shear Reynolds number based on the interfacial shear velocity uτI, and Tsat is the
saturation temperature at the local pressure.
Substitution the expression for Nu into the heat flux equation we obtain,
(2.42)
where C is the only model parameter. Note that this closed form expression is a special case due
the fact that the exponent of the Reynolds number in the Nusselt number expression is
1.
In case the velocities are zero, the interfacial shear phase change model will predict zero mass
transfer rate. The phase change in this case can occur due to thermal non-equilibrium which propagates
by diffusion. Two models are available for interfacial diffusion phase change, viz. the Schrage model and
the relaxation model.
For the Schrage model the heat transfer rate is given as,
(2.43)
where, Ac is the accommodation coefficient that is a material constant.
For the relaxation model (which is only valid for very diffusive problems) the heat transfer rate is
given as,
(2.44)
When the dispersed and interfacial models are both active, then the two models are blended based on
the magnitude of ∇α, where the interfacial model is chosen if the gradient is significant and
vice-versa.
Cavitation models
We start with the condition of vapour bubbles in liquid of radius RB with a nucleate density per unit
volume of liquid (n0). If we consider a volume Ω = ΩV + ΩL, then,
The bubble radius can be expressed using Eq. (2.45) as,
(2.47)
The other starting point for cavitation models is the simplified form of the Rayleigh-Plesset equation for
bubble growth given as,
(2.48)
The radius grows if p < pV and shrinks if p > pV. An expression for dαV∕dt using Eq. (2.45) is obtained
by assuming that n0 does not change with time,
(2.49)
The definition of mixture density (ρ = ρL + αV(ρV- ρL)) is used to simplify the mixture continuity
equation to obtain,
(2.50)
under the assumption of incompressibility. Further in the absence of slip between the vapour and liquid
phases (homogeneous flow), we can write the above equation as,
(2.51)
Under the assumptions of incompressibility and homogeneous flow the following expression for the
divergence of the velocity field is obtained,
(2.52)
where the subscript e denotes evaporation. Using Eqs. (2.51,2.52,2.49) we obtain an expression for the
evaporation rate as,
The problem with this model arises due to the fact that in a pure liquid αV = 0 which
makes the evaporation rate zero. Therefore an initial nucleation radius is introduced (R0)
and the related αV 0 can be calculated based on Eq. (2.46). The final model is written as,
The model has two input parameters n0 and R0, where typical values are n0 = 108 and R0 = 10-6 as in
Deimel et al. (2014).
using the notation of Singhal et al. (2002). In fact, with this correction the basic unadulterated
expression for the mass transfer rate (Singhal et al. (2002):Eq. (13)) is the same as Eq.
(2.53)
In this model the closure for RB is provided using critical Weber number considerations based on
stable bubble size under aerodynamic drag.
(2.60)
where Dd∞ is the bubble diameter at the end of the fragmentation process (stable diameter). Typically
a value between 5 < Wed∞< 20 is quoted in the literature (Kolev, 2002). By rearranging equation 2.60
for Dd∞, we get
In order to derive the exact model given in Singhal et al. (2002), we rearrange the equation as
follows,
(2.63)
Expressing the term in the square brackets as CVch∕σ, where C is a (dimensional with units of
velocity) empirical constant, and Vch is a characteristic velocity, we get,
(2.64)
The final form of the model is obtained by using different coefficients for the evaporation and
condensation rates. In addition, the volume fraction of vapour is removed from the evaporation rate
equation and the volume fraction of of liquid is removed from the condensation rate equation. This is
necessary since for a pure liquid αV is initially zero thereby making the evaporation rate zero.
Introducing these changes, we get
The volume fraction need to be converted to mass fraction in order to compare with Eqs. (15-16) in
Singhal et al. (2002). We obtain,
The model is closed by choosing Vch = and the velocities Ce = 0.02 and Cc = 0.01. The effect of
turbulence on the vapour pressure is accounted for by writing,
Implementation in TransAT
The evaporative mass transfer rate is written using an indicator function due to the change in
coefficients and the appropriate volume fraction as follows.
(2.73)
where the indicator function Ie is 1 if p < pV and 0 other wise and can be expressed as
(2.74)
2.2.2 Wall Heat Flux Modeling during Subcooled Nucleate Boiling
General Formulation
For the liquid phase, the heat flux from the wall to the liquid are transferred through the following
means: heat convection to the liquid phase away from the nucleation sites, microlayer evaporation and
quenching. In nucleate subcooled boiling, the wall heat is utilised in forming bubbles and heating the
liquid. According to Kurul & Podowski (1991), the total convective heat flux from the wall is the sum
of three models:
(2.75)
where q1ϕ′′ is the single-phase convective heat flux, qe′′ is the heat flux associated with phase change and
qQ′′ is the quenching flux, during the dwell time.
Single-Phase Heat Flux
In the portion of the wall unaffected by the bubbles, the heat transfer from the wall to the liquid can be
determined by:
(2.76)
Now, A1ϕ′′ is the fraction of the wall unaffected by the nucleation sites, ρl is the liquid density, cpl is the
liquid specific heat, Tw is the wall temperature, Tl is the temperature of the liquid. StP is the Stanton
Number calculated from heat transfer correlation in terms of the local liquid velocity and Prandtl
Number. StP could be written as,
(2.77)
where Nu, the Nusselt Number, is provided by the Ranz-Marshall correlation.
Re, Pr and Pe are the Reynolds, Prandtl and Peclet numbers respectively.
Evaporative Heat Flux
Taking into account the effect of area overlapping between the neighboring sites, the the evaporative
heat flux , qe′′ is given by E. Krepper & Egorov (2007), becomes
(2.80)
where ρv is the vapor density, hfg is the latent heat of vaporisation, ab2 is the bubble influence
factor having a value of 2. The area fraction under the influence of bubbles, A2ϕ′′ is given
by
(2.81)
where ddet is the bubble diameter at departure, which depends on the local liquid subcooling:
(2.82)
Here, ΔTsub is the difference between the saturation temperature and the liquid temperature.
(2.83)
In Eq. 2.80, fdet is the frequency of nucleation. In general, A simpler estimation of the bubble
departure frequency as ratio of the terminal rise velocity over the bubble departure diameter
is:
(2.84)
where g is the acceleration due to gravity. N′′ is the nucleation site density
(2.85)
Quenching Heat Flux
(2.86)
where hQ is the quenching heat transfer, given by
(2.87)
where, kl is the liquid thermal conductivity. Here, tw is the waiting time between the departure and the
inception of the next bubble. It is assumed that the waiting time takes 80% of the bubble departure
period. Hence,
(2.88)
To close the model, Eq. 2.75 can be solved iteratively for the local wall temperature Tw
2.3 Ensemble-averaged N-phase flow
2.3.1 N-phase model
A mixture of N phases can be simulated with the following set of equations. Definitions are given
below
with the subscript k denoting the phase index and the superscript m denoting a mixture variable, and α
being the volume fraction. Mixture variables are defined as follows
Note also that mixture viscosity can also be modelled. This will be detailed in the following
section.
2.3.2 Turbulence equations
For modelling the effect of turbulence for the N-phase flow, only the turbulent kinetic energy k and
dissipation rate ε of the mixture are solved for. Also, it is assumed that all the constants of the k - ε
model are the standard values for single phase flow. Equations for the turbulent variables can then be
written
2.4 Mixture viscosity
2.4.1 Suspension viscosity model
In a suspension, the flow is disturbed by the presence of a dispersed phase. If the concentration
of the dispersed phase increases, the particle or droplet will feel an increased resistance
of the flow, which results in a higher drag coefficient. This can be taken into account by
modifying the mixture viscosity (Ishii & Zuber, 1979; Hallanger et al., 1995; Manninen &
Taivassalo, 1996). In this section, subscript c denotes continuous phase, while d denotes dispersed
phase.
2.4.2 Suspended drops or bubbles
For a mixture including suspended drops or particles, Ishii & Zuber (1979) developed an apparent
viscosity expression, which reads
(2.100)
2.4.3 Solid suspensions
Several correlations are available for suspensions of solid particles. Ishii & Zuber (1979) propose the
following expression
(2.101)
with αCP is the close packing volume fraction, usually taken between 0.62 and 0.65 for solid
particles.
Finally, Mills (1985) derived the following equation for a suspension of hard spheres of equal
sizes
(2.103)
with αCP = .
Chapter 3 Particle Tracking Models
3.1 Mathematical formulation
3.1.1 Momentum equations
The equation governing the particle motion is the well known example given by Maxey & Riley (1983).
The equation is valid for a rigid sphere in non-uniform flow such that the sphere is isolated and far from
any boundary, the Reynolds number for the relative motion between the particle and the fluid is small,
and the particle size is much smaller than the smallest flow length scale. The equation is given as
follows,
where d∕dt denotes the time derivative along the particle path, and D∕Dt denotes the time derivative
following a fluid element. The terms in the right-hand side of Eq. (3.1) are referred to as the buoyancy,
fluid (due to the fluid pressure gradient and viscous stresses), added-mass, Stokes drag, and Basset
forces. For the case of particles much heavier than the fluid, Elghobashi & Truesdell (1992) have shown
that the only significant forces are the Stokes drag, buoyancy, and Basset forces. They also found that
the Basset force was always an order of magnitude smaller than the drag and buoyancy
forces.
Other forms of the forces and additional forces can be used under different circumstances.
Drag force
In the Stokes regime, drag force applied to the particle is written
(3.2)
For a higher Reynolds number based on particles, the Clift-Gauvin correlation can be used,
where
(3.3)
with a friction factor f being equal to
(3.4)
with the particle Reynolds being defined by
(3.5)
Lift force
The standard small particle (under the Stokes flow, Rep<< 1 condition) does not include a lift force
under the assumption of uniform flow upstream of the particle. However under non-uniform fluid flow
upstream of a particle a lift force exists as proposed by Saffman,
(3.6)
Auton (1987) derived the lift force acting on a sphere in shear flow under inviscid flow conditions
as,
(3.7)
For small particles, both of these lift forces are small when compared to the drag and
buoyancy forces. The lift force is also small if the dispersed phase is much heavier than the
fluid (solid particles or droplets in gas). However, for bubbly flow the lift force could be
significant.
Magnus Lift
The lift force due to particle rotation (Magnus force) is given for low Reynolds number by Rubinow &
Keller (1961) as,
(3.8)
where ωp is the particle rotation, and ∇× is the fluid rotation at the particle position. Rewriting it
in a general form with a lift coefficient CL,Mag we get,
(3.9)
where A is the projected area of the particle, and ωr = ωp-∇× is the relative rotation between
the particle and the fluid phase. The lift coefficient for low Reynolds number is then given
as
where γ, which is called the spin rate, is given as
(3.12)
where Rep is the particle Reynolds number.
Two-way Coupling
When two-way coupling is activated the forces acting on the particles also enter the Navier-Stokes
equations by means of an additional fluid-particle interaction force Ffp. For example the incompressible
Navier-Stokes equations become
where ρ represents the fluid density, gi is the acceleration of gravity, μ is the molecular viscosity and
σij is the viscous stress tensor of the fluid. ρ and μ are considered as constant for these
incompressible equations. Note that the Einstein summation convention applies to repeated indices.
Fifp is the component of the fluid-particle interaction force per unit volume in the cartesian direction
i,i = 1,2,3. As reported by Narayanan et al. (2002), for a given grid cell m, Fifp takes the following
form
(3.15)
where n is the particle index, Np is the total number of particles, Vp is the particle volume, ρf is the
fluid density, Vm is the volume of the finite-volume cell, fi is the component of the fluid-particle
interaction force in direction i on a single particle located at xn and W is the projection weight of the
force onto the cell node m located at xm.
3.1.2 Particle Angular momentum equation
(3.16)
where Ip is the moment of inertia, and Ti is the torque acting on the particle. The torque acting on a
spherical particle in Stokes flow is given as
(3.17)
For higher Reynolds numbers (up to 1000)
(3.18)
where Reω = ρfωpdp2∕4μf.
3.1.3 Temperature equation
Each particle is assumed to be represented by a uniform temperature Tp. This assumption is valid under
low particle Biot (Bip) number conditions, where Bip = Nupλ∕λp. The rate of change of particle
temperature is given as:
(3.19)
where, Nup is the particle Nusselt number given as
(3.20)
τpT is the particle thermal response time defined as
(3.21)
Tf(xp) is the fluid temperature at the particle position, Cp is the particle heat capacity, and λ is the
fluid thermal conductivity.
3.2 Particle Dispersion
This section describes the capabilities of TransAT for particle dispersion modelling using Langevin
equation. The particle dispersion model is used in combination with the Reynolds-Averaged turbulence
modelling approach.
3.2.1 Langevin model
The particle fluctuation variance (σp) is defined as a function of the fluid fluctuation variance (σf) and
the particle response time as:
(3.22)
where τf = MIN(τe,τr) is the effective eddy time scale, τp = ρpdp2∕18μ is the particle inertial response
time, K = 1 + 0.15Rep2∕3 is the drag correction for finite Reynolds number, and σf = . The model
constants Cτ and Cσ correspond to the input variables tauL_constant and sigmaF_constant,
respectively and are assigned default values of 0.11 and 1.69 ≈ 0.5, respectively (Lakehal
et al., 1995).
In τf, τe = Cτk∕ϵ is the local eddy life time in an Eulerian sense, and τr = Le∕|up-uf| is the time
taken by the particle to travel through the eddy (residence time), where Le = Cμ3∕4k3∕2∕ϵ is the
characteristic length scale of an eddy.
The particle velocity fluctuation equation is written as follows:
(3.23)
where τ′pi is the particle response time to turbulent fluctuations and ζ(t) is a random function with a
Gaussian probability density distribution and zero mean. The turbulent particle response time is given
as:
(3.24)
where uri is the terminal velocity due to action of body forces such as gravity.
3.3 Dense Particle Modelling
3.3.1 Eulerian-Lagrangian dense particle model
The Eulerian-Lagrangian formulation for dense particle flows featuring non-negligible volume fractions
(αP> 1%) in incompressible flow conditions is implemented in the CMFD code TransAT as follows
(mass and momentum equations for the fluid phase and Lagrangian particle equation of
motion):
(3.25)
(3.26)
(3.27)
(3.28)
(3.29)
where α is the in-cell volume of fluid (α + αP = 1), u is the velocity of the carrier phase, uP is the
particle velocity, u[xp] is the fluid velocity interpolated on to the particle position, ReP is the particle
Reynolds number, ΔUf-P is the relative velocity between the particle and fluid at the particle
position, Π is the sum of viscous stress σ and pressure p, τ is the turbulent stress tensor
(depending whether RANS, V-LES or LES is employed to deal with turbulence). Sources terms in
(3.26) denote body forces, Fb,and the rate of momentum exchange per volume between the
fluid and particle phases, FP. Various drag correlations are available in the literature to
account for higher volume fractions of particles; also the fluid viscosity can be modified based
on the volume fraction of particles. In (3.27), Fcoll denotes the inter-particle stress. The
momentum equation (3.26) presented here does not neglect viscous and turbulent diffusion
mechanisms in the fluid phase. The inter-phase drag model in (3.27) is set according to
Gydaspow (1989)
3.3.2 Four-way coupling modelling
The fluid-independent collision force Fcoll is made dependent on the gradient of the inter-particle stress,
π, using
(3.30)
Collisions between particles are estimated by the isotropic part of the inter-particle stress (its
off-diagonal elements are neglected.). π is modelled as a continuum stress (Harris & Crighton, 1994),
viz.
(3.31)
In TransAT, the particle field is predicted in a Lagrangian way first, then high-order accurate
interpolations are resorted to map the Eulerian field (to estimate π), then back again to the
Lagrangian system to determine Fcoll). The constant Ps has units of pressure, and αc is the
particle volume fraction at close packing (typically specified to be 0.6), and constant β is
set according to F.M. Auzerais & Russel (1988). The ε is a small number in the order of
10-7.
Collision/Contact Pressure Models
Snider Model
The collisional pressure model of Snider (2001) is defined as,
(3.32)
where Pp, β are model constants and ε = 10-7 is a small number that prevents the sign of the collision
pressure from changing for αp> αp,cp. Typical values for the model constants are given in Table 3.1.
where typical values for the constants are given in Table 3.2. The Jackson model is envisaged as an
add-on to kinetic models for granular flows and therefore has a cut-off particle volume fraction below
which it is set to zero. The Jackson model becomes qualitatively similar to the Snider model
(Snider, 2001), if αp,min = 0.
In order to apply the Jackson model in Lagrangian particle tracking, the change in sign of the
denominator for αp> αp,cp must be disallowed. Therefore, the Jackson model is implemented in
TransAT in a way similar to the Snider model.
The modelling of particle-boundary collision in TransAT is based on the hard sphere model described in
Crowe et al. (1998). This model is applied to interactions of particles with solid wall boundaries,
embedded objects and symmetry planes.
3.4.1 Hard sphere model
The hard sphere model for particle-wall collision is based on the integrated form of the momentum
equations of classical dynamics, where the deformation of the particle does not appear explicitly in the
formulation. The translational velocity p and angular velocity p after the collision is calculated from the
corresponding quantities before collision, the coefficient of restitution e and kinetic (sliding) friction f.
The deformation of the particle upon collision is assumed to be negligible. The particles are allowed to
slide along the wall, where Coulomb’s friction law is applied during this process. It is assumed that once
a particle has stopped sliding there is no further sliding.
The coefficient of restitution e is regarded as a property of the material and is the ratio of the normal
component of impulse in the compression period to that in the recovery period:
(3.35)
Coulomb’s law states that the friction force is the product of the normal force and the coefficient of
kinetic friction f. The friction force results in a change of the tangential component of impulse during
the sliding period according to
(3.36)
with corresponding to the unit vector in the direction of tangential velocity.
Note that the particle collision model is formulated for stationary boundaries. If the boundaries are
moving, the particle tangential velocities are transformed into a new, moving frame of reference, which
has the same velocity as the moving boundary, before the model equations are applied. Afterwards they
are transformed back to the original frame of reference.
3.4.2 Collision with wall boundary
Depending on whether the particle stops sliding during the collision process or continues throughout,
the translational and angular velocity of the particle after collision are calculated as follows (assuming a
plane wall with a normal vector in the positive y-direction): If vy0|vp|< -2_____
7f(e+1), then
vx
= 5
7(vx0-dp
5 ωz0)
(3.37)
vy
= -evy0
(3.38)
vz
= 5
7(vz0 + dp
5 ωx0)
(3.39)
ωx
= 2vzdp
(3.40)
ωy
= ωy0
(3.41)
ωz
= -2vxdp
(3.42)
If -2_____
7f(e+1)<vy0|vp|< 0, then
vx
= vx0 + εxf(e + 1)vy0
(3.43)
vy
= -evy0
(3.44)
vz
= vz0 + εzf(e + 1)vy0
(3.45)
ωx
= ωx0- 5 _
dpεzf(e + 1)vy0
(3.46)
ωy
= ωy0
(3.47)
ωz
= ωz0- 5 _
dpεxf(e + 1)vy0
(3.48)
The superscript 0 denotes the initial state before collision and εx,z corresponds to the components of
the tangential velocity unit vector in the respective direction, i.e. they satisfy the following
equation.
(3.49)
3.4.3 Collision with symmetry boundary
When particles collide with a symmetry boundary the collision model coefficients of restitution and
friction take the following values
e = 1.0
f = 0.0
(3.50)
In this case the denominator of the condition for distinction becomes zero and the following relation
holds for all non-zero particle velocities.
(3.51)
The corresponding equations for the translational and rotational velocities after collision simplify
to
vx
= vx0
(3.52)
vy
= -vy0
(3.53)
vz
= vz0
(3.54)
ωx
= ωx0
(3.55)
ωy
= ωy0
(3.56)
ωz
= ωz0
(3.57)
i.e. only the sign of the wall normal velocity component is changed, the remaining translational
velocity components and the rotational velocities do not change, thus the total translational and
rotational momentum is conserved.
Chapter 4 Hydrates Modelling
Hydrates modelling is available in TransAT in conjunction with the N-phase mixture model (See
Chapter Phase-Average Mixture Models).
4.1 Hydrate Formation Kinetics
Two models have been implemented for hydrate formation: finite–rate kinetics model (Vysniauskas &
Bishnoi, 1983) and instantaneous formation models as formation kinetics can be considered to be
negligible compared to the other time scales of the flow.
4.1.1 Finite Rate Kinetics
A finite-rate kinetics model for the formation of hydrates (based on interfacial super-saturation) was
implemented in TransAT. It is called the Bishnoi model. This model (Vysniauskas & Bishnoi, 1983) is as
follows:
(4.1)
with the default values used for the constants listed in Table 4.1.
The instantaneous formation model has been implemented in TransAT as follows.
Similar to the finite-rate model, the temperature in each cell is compared to the Teq
for formation of hydrates. Several curves are available in TransAT for the following
materials:
Methane
Ethane
Propane
Isobutane
If the temperature in the computational cell is higher than Teq then no hydrates are
formed.
Note that if the hydrate formation rate is high, then locally more heat will be released resulting in
a higher temperature. If the temperature goes higher than Teq then hydrate formation is inhibited.
The iterations per time step in the implicit method allow one to obtain the converged solution of
the correct temperature and reaction rate.
The number of moles of hydrate forming phase (hf) in a computational cell is given
as:
(4.2)
where, Mwhf is the molecular weight of hydrate forming phase, Mhf is the number of moles
of the hydrate forming phase in the cell, and Ωi is the volume of the computational
cell.
The number of moles of water in the computational cell is given as:
(4.3)
where Mww is the molecular weight of water, and Mw is the number of moles of water available
for hydrate formation.
The rate of formation of hydrates ṙ is then given as:
(4.4)
where nhy is the hydration number of the hydrate forming phase. The reaction rate is calculated
such that all the hydrate forming phase present in a cell is consumed within one time step
(provided that the temperature is less than Teq).
The hydrate formation rate ṙ is updated with a relaxation factor βhy in every iteration for a given
time step as follows.
(4.5)
where (k + 1) denotes the current iteration, and (k) denotes the previous iteration. Relaxation
is a measure to prevent oscillations in the reaction rate which in turn prevents the
convergence of the volume fraction equations. A relaxation factor of 0.4 is used in
TransAT.
4.2 Hydrate Dissociation and Melting
4.2.1 Instantaneous Hydrates Dissociation
An instantaneous hydrates dissociation model has been implemented in TransAT.
The temperature in each cell is compared to the Teq for formation of hydrates. The same
curve as for hydrate formation is used. If the temperature in the computational cell is lower
than Teq by ΔTsh then hydrates can form, where ΔTsh is the equilibrium range allowed.
If the temperature in the computational cell is higher than Teq then hydrates dissociate.
Note that if the hydrate dissociation rate is high, then locally more heat will be absorbed
resulting in a lower temperature. If the temperature goes lower than Teq then hydrate
dissociation is inhibited. The iterations per time step in the implicit method, allow one to
obtain the converged solution of the correct temperature and reaction rate.
The number of moles of hydrates (Mhy) in a computational cell is given as:
(4.6)
where, Mwhy is the molecular weight of the hydrate phase, Mhy is the number of moles of the
hydrate forming phase in the cell, and Ωi is the volume of the computational cell. The rate of
dissociation of hydrates ṙ is then given as:
(4.7)
The reaction rate is calculated such that all the hydrate phase present in a cell is consumed within
one time step, provided that the temperature is higher than Teq + ΔTsh.
The hydrate dissociation rate ṙ is updated with a relaxation factor βhy in every iteration for a
given time step as follows.
(4.8)
where (k + 1) denotes the current iteration, and (k) denotes the previous iteration. Relaxation is a
measure to prevent oscillations in the reaction rate A relaxation factor of 0.4 is used in
TransAT.
4.2.2 Heat Absorption
The formation of hydrates is exothermic. The dissociation of hydrates into water and gas phases is
endothermic and results in absorption of heat locally. The enthalpy of dissociation is taken to be the
same as the enthalpy of formation.
The heat absorbed is included in the mixture temperature equation as a sink term given
by,
(4.9)
where Hhy,diss is the enthalpy of hydrate dissociation, and ṙdiss is the rate of dissociation
of hydrates. With the dissociation and the related heat absorption, the hydrate volume
fraction at any location is a non-linear dynamic balance between possible formation and
dissociation (with a gap of Tsh- Tsc between the formation and dissociation limits for numerical
stability).
4.3 Hydrate Rheology
A special hydrate rheology model has been implemented in TransAT.
First, the continuous phase is defined depending on the Water Cut. If it is more than 50%, water is
the continuous phase. Otherwise, oil is the continuous phase.
The volume fraction of the dispersed phase is then calculated as follows
(4.10)
where αwater in hydrates = Hαhydrates, with H the hydration number. The viscosity of the slurry is
then calculated using Mills law (Mills, 1985):
(4.11)
with
(4.12)
where τ denotes the instantaneous shear stress. For numerical reasons, in the implementation τ0 cannot
be smaller than 10-6 Pa and αeff must not be larger than 0.997αmax.
Bibliography
Antal, S. P., Lahey, R. T. & Flaherty, J. E. 1991 Analysis of phase distribution in
fully developed laminar bubbly two-phase flow. Int. J. Multiphase Flow 17 (5), 635–652.
Auton, T. R. 1987 The lift force on a spherical body in a rotational flow. J. Fluid Mech.183, 199–218.
Banerjee, S., Lakehal, D. & Fulgosi, M. 2004 Surface divergence models for scalar
exchange between turbulent streams. International Journal of Multiphase Flow 30, 963–977.
Banerjee, S., Rhodes, S. & Scott, D. 1968 Mass transfer through falling wavy liquid
films in turbulent flows. Industrial & Engineering Chemistry Fundamentals 7, 22.
Coste, P. 2004 Computational simulation of multi-D liquid-vapor thermal shock with
condensation. In Proceedings of ICMF’04. Yokohama, Japan.
Crowe, C., Sommerfeld, M. & Tsuji, Y. 1998 Multiphase flows with droplets andparticles. Boca Raton, Florida, USA: CRC Press.
Davis, J. & Yadigaroglu, G. 2004 Direct contact condensation in hiemenz flow boundary
layers. Int. J. Heat and Mass Transfer 47, 1863–1875.
De Gennes, P.-G. 1985 Wetting: statics and dynamics. Reviews of modern physics 57 (3),
827.
Deimel, C., Günther, M. & Skoda, R. 2014 Application of a pressure based cfd code
with mass transfer model based on the rayleigh equation for the numerical simulation of the
cavitating flow around a hydrofoil with circular leading edge. In EPJ Web of Conferences, ,
vol. 67, p. 02018. EDP Sciences.
Drew, D. A. & Lahey, R. T. 1987 The virtual mass and lift force on a sphere in rotating
and straining nviscid flow. Int. J. Multiphase Flow 13, 113–121.
Drew, D. A. & Passman, S. L. 1998 Theory of multicomponent fluids. Springer-Verlag
New York, Inc.
E. Krepper, B. K. & Egorov, Y. 2007 Cfd modelling of subcooled boiling concept,
validation and application to fuel assembly design. Nuclear Engineering and Design 237,
716–731.
Elghobashi, S. & Truesdell, G. C. 1992 Direct simulation of particle dispersion in a
decaying isotropic turbulence. J. Fluid Mech. 242, 655–700.
F.M. Auzerais, R. J. & Russel, W. 1988 The resolution of shocks and the effects of
compresisble sediments in transient settling. J. Fluid Mech. 195, 437.
Frank, T., Shi, J. M. & Burns, A. D. 2004 Validation of eulerian multiphase flow models
for nuclear safety applications. In 3rd International Symposium on Two-Phase Flow Modellingand Experimentation. Pisa, Italy.
Friess, H. & Lakehal, D. 2004 A new method for including interfacial tension and wetting
dynamics in the simulation of two-phase flow. In ICHMT DIGITAL LIBRARY ONLINE. Begel
House Inc.
Gydaspow, D. 1989 Hydrodymanmics of fluidization and heat transfer supercomputer
mmodeling. Appl. Mech. Rev. 39, 1.
Hallanger, A., Sveberg, K., Anfinsen, N., Sènstabè, I. & Knutsen, T. 1995 A
simulation model for three-phase gravity separators. In 2nd Colloquiun on Process Simulation,
pp. 199–213. Helsinki University of Technology, Espoo, Finland.
Harris, S. E. & Crighton, D. G. 1994 Solutions, solitary waves and voidage disturbances
in gas-fluidized beds. J. Fluid Mech. 266, 243.
Hosokawa, S., Tomiyama, A., Misaki, S. & Hamada, T. 2002 Lateral migration of single
bubbles due to the presence of wall. In In Prococeedings of ASME Joint U.S.European FluidsEngineering Division Conference. Montreal, Canada.
Hughes, E. & Duffey, R. 1991 Direct contact condensation and momentum transfer in
turbulent separated flows. International Journal of Multiphase Flow 17, 599–619.
Ishii, M. 1975 Thermo-Fluid Dynamic Theory of Two-Phase Flow. Paris, France: Eyrolles.
Ishii, M. & Mishima, K. 1984 Two-fluid model and hydrodynamic constitutive relations.
Nucl. Eng. and Des. 82, 107–126.
Ishii, M. & Zuber, N. 1979 Drag coefficient and relative velocity in bubbly, droplet or
particulate lows. AIChE Journal 25, 843.
Johnson, P. C. & Jackson, R. 1987 Frictional-collisional constitutive relations for granular
materials, with application to plane shearing. J. Fluid Mech. 176, 64–93.
Kolev, N. 2002 Multiphase Flow Dynamics: Mechanical and thermal interactions.
Engineering online library . Springer.
Kurul, N. & Podowski, M. Z. 1991 On the modeling of multidimensional effects in boiling
channels. Proceedings of the 27th National Heat Transfer Conference pp. 30–40.
Lakehal, D., Fulgosi, M., Banerjee, S. & Yadigaroglu, G. 2008a Turbulence and
heat exchange in condensing vapor-liquid flow. Physics of Fluids 20, 065101.
Lakehal, D., Fulgosi, M. & Yadigaroglu, G. 2008b DNS of a condensing stratified
steam-water flow. Journal of Heat Transfer 130, 02501–10.
Lakehal, D. & Labois, M. 2011a A new modelling strategy for phase-change heat transfer
in turbulent interfacial two-phase flow. Int. J. Multiphase Flow 37, 627–639.
Lakehal, D. & Labois, M. 2011b A new modelling strategy for phase-change heat transfer
in turbulent interfacial two-phase flow. International Journal of Multiphase Flow 37 (6),
627–639.
Lakehal, D., Mestayer, P. G., Edson, J. B., Anquetin, S. & Sini, J.-F. 1995
Eulero-lagrangian simulation of raindrop trajectories and impacts within the urban canopy.
Atmospheric Environment 29, 3501–3517.
Lim, I. S., Tankin, R. S. & Yuen, M. C. 1984 Condensation measurement of horizontal
cocurrent steam-water flow. ASME J. Heat Transfer 106, 425–432.
Manninen, M. & Taivassalo, V. 1996 On the mixture model for multiphase flow. Espoo,
Finland: VTT Publications.
Maxey, M. R. & Riley, J. K. 1983 Equation of motion for a small rigid sphere in a
nonuniform flow. Phys. Fluids 26, 883–889.
Mills, P. 1985 Non-newtonian behaviour of flocculated suspensions. Journal de Physique –Lettres 46, L301–L309.
Narayanan, C. & Lakehal, D. 2006 Simulation of filling of microfluidic devices using a
coarse-grained continuum contact angle model. In NSTI Nanotechnology Conference and TradeShow, , vol. 711. Boston.
Narayanan, C., Lakehal, D. & Yadigaroglu, G. 2002 Linear stability analysis of
particle–laden mixing layers using lagrangian particle tracking. Powder Tech. 125, 122–130.
Oesterl, B. & Dinh, T. B. 1998 Experiments on the lift of a spinning sphere in a range
of intermediate reynolds numbers. Experiments in Fluids 25, 16–22, 10.1007/s003480050203.
Osher, S. & Sethian, J. A. 1988 Fronts propagating with curvature-dependent speed:
Algorithms based on hamilton-jacobi formulations. J. Comp. Phys. 79, 12–49.
Qian, T., Wang, X.-P. & Sheng, P. 2004 Power-law slip profile of the moving contact line
in two-phase immiscible flows. Physical review letters 93 (9), 094501.
Ranz, W. E. & Marshall, W. R. 1952 Evaporation from drops. Chemical EngineeringProgress 48 (3), 141–146.
Rubinow, S. I. & Keller, J. B. 1961 The transverse force on a spinning sphere moving
in a viscous fluid. Journal of Fluid Mechanics 11 (03), 447–459.
Rzehak, R., Krepper, E. & Lifante, C. 2012 Comparative study of wall-force models for
the simulation of bubbly flows. Nucl. Eng. Des. 253 (0), 41–49.
Schiller, L. & Naumann, A. Z. 1933 A drag coefficient correlation. Ver. Deutch Ing.77 (0), 318–320.
Singhal, A. K., Athavale, M. M., Li, H. & Jiang, Y. 2002 Mathematical basis and
validation of the full cavitation model. Journal of fluids engineering 124 (3), 617–624.
Snider, D. M. 2001 An incompressible three-dimensional multiphase particle-in-cell model
for dense particle flows. Journal of Computational Physics 170, 523–549.
Sussman, M., Smereka, P. & Osher, S. 1994 A level set approach for computing solutions
to incompressible two-phase flow. J. Comp. Phys. 114, 146–159.
Theophanous, T., Houze, R. & Brumfield, L. 1976 Turbulent mass transfer at free
gas-liquid interfaces, with application to open-channel, bubble and jet flow. InternationalJournal of Multiphase Flow 19, 613–623.
Tomiyama, A. 1998a Drag coefficients of single bubbles under normal and micro gravity
conditions 41, 472–479.
Tomiyama, A. 1998b Struggle with computational bubble dynamics. In Proc. of the ThirdInternational Conference on Multiphase Flow. ICMF’98, Lyon, France.
Tomiyama, A., Sou, A., Zun, I., Kanami, N. & Sakaguchi, T. 1995 Effects of eotvos
number and dimensionless liquid volumetric flux on lateral motion of a bubble in a laminar
duct flow. In Proceedings of the Second International Conference on Multiphase Flow, pp. 3–16.
Kyoto, Japan.
Ubbink, O. 1997 Numerical prediction of two fluid systems with sharp interfaces. Doctoral
thesis, Imperial College London (University of London).
Ubbink, O. & Issa, R. 1999 A method for capturing sharp fluid interfaces on arbitrary
meshes. Journal of computational physics 153 (1), 26–50.
Vyskocil, L. & Macek, J. 2008 Boiling flow simulation in neptune cfd and fluent codes.
XCFD4NRS .
Vysniauskas, A. & Bishnoi, P. R. 1983 A kinetic study of methane hydrate formation.
Chemical Engineering Science 38 (7), 1061–1072.
Yuan, W., Sauer, J. & Schnerr, G. H. 2001 Modeling and computation of unsteady
cavitation flows in injection nozzles. Mécanique & industries 2 (5), 383–394.