Algorithm 1
Algorithm 1
DOI: 10.1002/htj.21268
RESEARCH ARTICLE
Nomenclature: 𝐶𝛼 , Compression factor (−); 𝐶𝑝 , Specific heat (J/kgK); 𝑑𝐿 , Liquid thermal diffusivity: 𝑘∕𝜌𝐶 (m2 /s); 𝐷, Equiv-
𝑝
alent diameter (M); 𝐷0 , Bubble Initial diameter (M); 𝑔,⃗ gravity acceleration (m/s2 ); 𝐻𝐿𝐺 , Latent heat (J/kg); 𝑘, Thermal con-
ductivity (W/mK); 𝑀, Molar mass (kg/Kmol); ṁ , mass flow rate per unit area (kg/m2 s); ṁ ′′′ , mass flow rate per unit volume
′′
(kg/m3 s); 𝑃 , Pressure (Pa); 𝑃𝑟𝑔ℎ , Dynamic Pressure (Pa); 𝑟, Mass transfer intensity factor (1/s); 𝑅, Universal gas constant (J/mol
K); 𝑇 , Temperature (K); 𝑊 , width (M); 𝑈⃗ , velocity (m/s); 𝑈⃗𝑐 , Compressive velocity (m/s)
Greek symbols: 𝛼, Volume fraction factor; 𝜅, Interface curvature (m−1 ); 𝜇, Dynamic viscosity (Pa.s); 𝜆, Taylor unstable wave
(M); 𝜌, Density (kg/m3 ); 𝛾, Tanasawa Coefficient (−); 𝜎, Surface Tension (N/m)
Subscripts: 𝑐, Condensation; 𝑒, Evaporation; 𝐺, Gas (vapor) phase; 𝐿, Liquid phase; sat, Saturation; sub, Subcooled; sup,
Superheat; W, wall
Heat Transfer—Asian Res 2017; 00: 1–31 [Link]/journal/htj © 2017 Wiley Periodicals, Inc. 1
2 SAMKHANIANI AND ANSARI
KEYWORDS
boiling, condensation, interFoam, OpenFOAM, volume of fluid
1 I N T RO D U C T I O N
and physics involving extensive topology changes. Therefore, volume of fluid method without interface
reconstruction which called color function volume of fluid has gained more popularity.
The Eulerian approach has been developed to simulate flow with phase change in the last decades.
Welch and Wilson16 introduced a PLIC-VOF approach for simulation of boiling. Hardt and Wondra17
proposed an evaporation model that relies on a continuum field representation of the source terms in
the mass conservation equation and applied on PLIC-VOF in Fluent software. In order to simulate
evaporating droplets, Schlottke and Weigand18 developed a PLIC-VOF method which involves two
VOF variables, one for the Liquid phase and one for the vapor phase. Son and Dhir19 and Son et al.20
modified the Level set method to account for phase change. They assumed that the vapor phase is at
saturation temperature, so they solved the energy equation only for the liquid phase. Level set methods
for incompressible flow with phase change are also presented by Tanguy et al.21 and Gibou et al.22 who
applied the ghost fluid approach to impose the jump conditions at the interface. A CLSVOF method for
simulation of bubble growth in film boiling was developed by Tomar et al.23 Guo et al.64 simulated film
boiling using VOSET method. Kunkelmann and Stephan24 simulated nucleating boiling with special
treatment for heat flux in contact angle.
There are several Eulerian codes to simulate multiphase flows, a color function volume of fluid
(CF-VOF) solver named interFoam in open source CFD packge OpenFOAM25 has been receiving
an increased amount of attention and usage.26–28 This VOF solver was implemented by Ubbink29
at first, then it has undergone several modifications. The present code is a part of C++ libraries of
OpenFOAM. This software was designed for finite volume discretization of a generic set of partial
differential equations. However, several useful features such as ease of parallelization, availability
of preprocessing and post processing utilities, error checks and choice of spatial and temporal dis-
cretization schemes make the code attractive for research. InterFoam has been recently contributed to
consider phase change phenomena (boiling and condensation). Table 1 provides a comprehensive sum-
mery of literature concerning the implementation of different computational phase change model using
OpenFOAM.
4 SAMKHANIANI AND ANSARI
√ [ ]
′′ 2 𝑀 𝑃𝐺 𝑃𝐿
𝑚̇ = 𝛾 √ − 𝛾𝑒 √ , (1)
2 − 𝛾𝑐 2𝜋𝑅 𝑐 𝑇𝐺 𝑇𝐿
where 𝑅 = 8.314 J/mol K is the universal gas constant, 𝛾 the fraction of molecules transferred from one
phase to the other during phase change. The subscripts 𝑐 and 𝑒 in the equation refer to condensation
and evaporation, respectively. Here𝛾𝑒 = 1 and 𝛾𝑐 represent complete evaporation and complete con-
densation, respectively. In phase change simulation, usually equal values of 𝛾𝑐 and 𝛾𝑒 are considered.
While 𝛾 = 0.1 − 1 for dynamically renewing water surfaces such as jets and moving films and 𝛾 < 0.1
for stagnant surfaces are recommended.37 Tanasawa38 simplified Eq. (1) using extra assumptions. He
assumed that the interface is at saturation temperature and the heat flux linearly dependent on the tem-
perature jump between the interface and the vapor. The mass flux rate in Tanasawa’s model is given by
√
′′2𝛾 𝑀 𝜌𝐺 𝐻𝐿𝐺 (𝑇 − 𝑇𝑠𝑎𝑡 )
𝑚̇ = , (2)
2−𝛾 2𝜋𝑅 𝑇
3∕2
𝑠𝑎𝑡
where 𝑇𝑠𝑎𝑡 is based on local pressure. The volumetric transferred mass rate in (kg/m3 s) is defined as
′′′ ′′
𝑚̇ = 𝑚̇ |∇𝛼𝐿 |. (3)
SAMKHANIANI AND ANSARI 5
′′′ 𝑇 − 𝑇𝑠𝑎𝑡
𝑚̇ = 𝑟𝑐 (1 − 𝛼𝐿 )𝜌𝐺 , for condensation 𝑇 < 𝑇𝑠𝑎𝑡 (4)
𝑇𝑠𝑎𝑡
and
′′′ 𝑇 − 𝑇𝑠𝑎𝑡
𝑚̇ = 𝑟𝑒 𝛼𝐿 𝜌𝐿 , for evaporation 𝑇 > 𝑇𝑠𝑎𝑡 , (5)
𝑇𝑠𝑎𝑡
where 𝑟𝑐 and 𝑟𝑒 in 𝑠−1 are empirical coefficients called the mass transfer intensity factor. The extensive
ranges from 0.1 to 107 have been conducted in literature for 𝑟.40–45
where 𝑘𝑒𝑓 𝑓 is the effective thermal conductivity, which can be determined from the volume fractions
and thermal conductivities of the vapor and liquid.46
2. COMPUTAT I O NA L ME T H O D
fixed grids. This scalar function is the ratio of one fluid volume to the volume of cell, and it is defined
as
⎧1 𝑥⃖⃗ ∈ Liquid
𝑉𝐿𝑖𝑞𝑢𝑖𝑑 ⎪
𝛼𝐿 (⃖𝑥,
⃗ t) = = ⎨0 < 𝛼𝐿 < 1 𝑥⃖⃗ ∈ interface . (7)
𝑉 ⎪
⎩0 𝑥⃖⃗ ∈ vapor
The thermo-physical properties of two-phase flow such as viscosity (𝜇), density (𝜌), and thermal
conductivity (𝑘) are defined as
where 𝑦 has liquid physical properties where 𝛼𝐿 = 1.0, and it has vapor physical properties where
𝛼𝐿 = 0.0. The interface is moving along the flow; therefore the volume fraction is advected to keep
the interface. In order to derive the transport equation for volume fraction, density (𝜌) from Eq. (8) is
placed in Eq. (9) and rearranged to obtain Eq. (10). Eq. (9) indicates the global continuity equation
𝜕 ⃖⃖⃗) = 0,
(𝜌) + ∇.(𝜌𝑈 (9)
𝜕𝑡
𝜕𝛼𝐿 𝜌 ∇.𝑈⃖⃖⃗
+𝑈 ⃖⃖⃗ = − 𝐺
⃖⃖⃗.∇𝛼𝐿 + 𝛼𝐿 ∇.𝑈 , (10)
𝜕𝑡 (𝜌𝐿 − 𝜌𝐺 )
where ∇.𝑈⃖⃖⃗ is zero in an incompressible adiabatic two-phase flow. In the phase change process, it is
determined from the local continuity for each phase:
⃖⃖⃗.⃖⃖⃖ 𝑚̇
𝑈 𝑛𝐿⃗𝑑𝑆𝐿 = − , (13)
∫ 𝜌𝐿
𝐿
⃖⃖⃗.⃖⃖⃖⃖ 𝑚̇
𝑈 𝑛𝐺⃗𝑑𝑆𝐺 = + . (14)
∫ 𝜌𝐺
𝐺
The unit normal vector at the interface is displayed in Fig. 2. As 𝑛⃖⃗𝐺 = 𝑛⃖⃗𝐼 = −⃖𝑛⃗𝐿 , then the summation
of the former equations results in
( )
⃖⃖⃗.⃖𝑛⃗𝑑𝑆 = 𝑚̇ 1 1
𝑈 − . (15)
∫ 𝜌𝐺 𝜌𝐿
𝑆
SAMKHANIANI AND ANSARI 7
FIGURE 2 Two-phase flow system, S denotes surface and n denotes unit normal, L: liquid, G: vapor, and
I: interface
By replacing Eq. (16) into Eq. (10), the following transport equation for 𝛼𝐿 is derived:
[ ( )]
𝜕𝛼𝐿 1 1 1
⃖⃖
⃗
+ 𝑈 .∇𝛼𝐿 = −𝑚̇
′′′
− 𝛼𝐿 − . (17)
𝜕𝑡 𝜌𝐿 𝜌𝐿 𝜌𝐺
In OpenFOAM, the extra divergence term is added to Eq. (17). This term contributes only in the
region of the interface (0 < 𝛼𝐿 < 1.0). It limits the smearing of the interface because of the compen-
sation of the diffusive fluxes28
[ ( )]
𝜕𝛼𝐿 1 1 1
⃖⃖
⃗ ⃖⃖⃖⃗
+ 𝑈 .∇𝛼𝐿 + ∇.(𝛼𝐿 (1 − 𝛼𝐿 )𝑈𝑐 ) = −𝑚̇
′′′
− 𝛼𝐿 − . (18)
𝜕𝑡 𝜌𝐿 𝜌𝐿 𝜌𝐺
⃖⃖⃖⃗𝑐 is compressive velocity. It is calculated in the normal direction to the interface to avoid any
Here 𝑈
dispersion.47 It is defined as
∇𝛼𝐿
⃖⃖⃖⃗𝑐 = min{𝐶𝛼 |𝑈 |, max(|𝑈 |)}
𝑈 . (19)
|∇𝛼𝐿 |
The coefficient 𝐶𝛼 controls the weight of the compression flux and should usually be in the range
of unity (1.0 < 𝐶𝛼 < 4.0).28,48,49 In the present study, the compression factor 𝐶𝛼 = 1.0 is applied. The
large compression factor increases the non-physical spurious current in the vicinity of the interface.
The momentum equation is given by
⃖⃖⃗)
𝜕(𝜌𝑈 ⃖⃖⃗𝑈
+ ∇.(𝜌𝑈 ⃖⃖⃗) − ∇.(𝜇(∇𝑈 ⃖⃖⃗)) = −∇𝑃 + 𝜌⃖𝑔⃗ + 𝜎𝜅∇𝛼𝐿 .
⃖⃖⃗𝑇 + ∇𝑈 (20)
𝜕𝑡
The last term in the right hand of Eq. (20) is considering surface tension using the continuous surface
force (CSF) model.50 Where 𝜎 is the surface tension, and 𝜅 is the interface curvature and is defined as
( )
𝛼𝐿
∇̃
𝜅 = −∇. , (21)
|∇̃
𝛼𝐿 |
8 SAMKHANIANI AND ANSARI
where 𝛼̃𝐿 is a smoothed VOF function calculated from 𝛼𝐿 by smoothing it over a finite region around the
interface. In the VOF method, the fluid interface sharply changes over a thin region. This abrupt change
of the VOF function creates errors in calculating the normal vectors and the curvature of the interface,
which are used to evaluate the interfacial forces. These errors induce spurious currents in the interfacial
region.28,51 The spurious currents create extra heat convection around the interface which increases
local mass transfer. An easy way to suppress these artifacts is to compute the interface curvature from
a smoothed VOF function. In the present study, the smoother proposed by Lafaurie et al.52 is applied:
∑𝑛
𝑓 =1 𝛼𝐿𝑓 𝑆𝑓
𝛼
̃𝐿𝑝 = ∑𝑛 , (22)
𝑓 =1 𝑆𝑓
where 𝑆𝑓 is the magnitude of face area, the subscript 𝑝 denotes the cell index, and 𝑓 denotes the
face index. The interpolated value (𝛼𝐿𝑓 ) at the face center is calculated using linear interpolation.
The application of this filter can be repeated 𝑚 times to get a smoothed field. It should be noted that
smoothing tends to wipe out high curvature regions and should be applied only up to the level that is
strictly necessary to sufficiently suppress parasitic currents. Hoang et al.53 found that the magnitude
of parasitic current decreases up to one order from 𝑚 = 0 to 𝑚 = 2 and only a slight further decrease
was observed for 𝑚 > 2. Therefore, in present study 𝑚 = 2 is employed in all simulations.
Energy equation is given by
𝜕 ′′′
(𝜌𝐶𝑝 𝑇 ) + ∇.(𝜌𝐶𝑝 𝑈 𝑇 ) − ∇.(𝑘∇𝑇 ) = −𝑚̇ 𝐻𝐿𝐺 , (23)
𝜕𝑡
where 𝐻𝐿𝐺 is latent heat coefficient. The last term in Eq. (23) is added to consider latent heat during
phase change.
The interface temperature is often at saturation, it is almost constant during phase change. However,
the slight variation in the quantity of saturation temperature as a function of local pressure is determined
using a simplified version of the Clausius–Clapeyron relation which is based on ideal gas assumption
for the vapor phase
( )
𝑃𝑠𝑎𝑡,1 𝑀𝐻𝐿𝐺 1 1
ln =− − . (24)
𝑃𝑠𝑎𝑡,0 𝑅 𝑇𝑠𝑎𝑡,1 𝑇𝑠𝑎𝑡,0
It is well known that numerical interfacial flow computations become more challenging when the
imbalance of material properties between the phases increases.51 In order to reduce the material prop-
erties imbalance across the interface, and increase numerical robustness, Eq. (23) is rearranged as
𝜕 ′′′
(𝑇 ) + ∇.(𝑈 𝑇 ) − ∇.(𝐷𝑘 ∇𝑇 ) = −𝐷𝑐 𝑚̇ 𝐻𝐿𝐺 , (25)
𝜕𝑡
where 𝐷𝑘 and 𝐷𝑐 are defined as
1.0
𝐷𝑐 = , (26)
𝜌𝐿 𝐶𝐿 𝛼𝐿 + 𝜌𝐺 𝐶𝐺 (1.0 − 𝛼𝐿 )
𝑘𝐿 𝛼𝐿 + 𝑘𝐺 (1.0 − 𝛼𝐿 )
𝐷𝑘 = . (27)
𝜌𝐿 𝐶𝐿 𝛼𝐿 + 𝜌𝐺 𝐶𝐺 (1.0 − 𝛼𝐿 )
To close the above set of equations, an appropriate phase change model is necessary to calculate
′′′
transferred mass flux rate (𝑚̇ ). In the present study, Lee and Tanasawa mass transfer model are
employed.
SAMKHANIANI AND ANSARI 9
(PISO).55 In the OpenFOAM frame work, PISO can be repeated for multiple iterations at each time
step. This process is referred to as the PIMPLE loop. Here, only one outer correction is applied and the
PIMPLE loop works in PISO mode. First, the matrix equation for the momentum equation is formed.
Then the inner pressure–velocity correction process is initiated. In PISO, an intermediate velocity
field is first obtained, and the cell-face volume fluxes (𝜙) are evaluated and corrected for gravitational
forces, the continuum surface-tension force, and boundary conditions. The pressure-Poisson equation
is then formed and solved. Following the approach of Ref. 25 the coefficients of the pressure equation
are obtained from the diagonal entries of the momentum matrix equation (𝐴𝐷 ). For condensing and
evaporating flow, the pressure equation would be
( ) ( )
1 ′′′ 1 1
∇. ∇𝑝 = ∇.𝜙 − 𝑚̇ − . (28)
𝐴𝐷 𝜌𝐺 𝜌𝐿
The last term on the left hand of Eq. (28) is considering the phase change process. The modified pres-
sure equation (pEqn.H) is shown in Fig. 5. In order to avoid a checker-boarding effect in the momentum
equation, the Rhie–Chow momentum interpolation56 is applied.
Finally, the thermal energy transport equation (Eq. (25)) is solved. The TEqn.H is shown in Fig. 6.
For numerical stability, source terms in TEqn.H and pEqn.H are linearized.
The terms vDotAlpha, vDotP and vDotT are mass flux rates which are calculated based on selected
mass transfer models. Lee mass transform model implemented in Lee.C is shown in Fig. 7.
The discretization scheme is selectable by the user; in this series of simulations the schemes dis-
played in Table 2 are chosen. For convenience, the corresponding terminology of OpenFOAM is given.
12 SAMKHANIANI AND ANSARI
3. VA L I DAT I O N A N D V E R I F I CAT I O N
In this section, for validation of the present solver, five problems are considered. Among them, the
Stefan problem, horizontal film condensation on a horizontal plate, and laminar film condensation on
a vertical plate have analytical solutions. For the others including 2D film boiling and bubble conden-
sations, the CFD result is compared with experimental data and correlations or previous numerical
simulations using the sharp interface method. Since there is no interface reconstruction in the present
VOF method, an iso-contour 𝛼𝐿 = 0.5 is assumed as the interface in all following simulations.
equilibrium. Evaporation pushes liquid away from the heated wall. The analytical solution for interface
position 𝛿(𝑡) and temperature 𝑇 (𝑥, 𝑡) on the vapor side are given by16
√
𝛿𝑎𝑛 (𝑡) = 2𝜂 𝑑𝐺 𝑡, (29)
( )
𝑇 − 𝑇𝑤 𝑥
𝑇 (𝑥, 𝑡) = 𝑇𝑤 + 𝑠𝑎𝑡 𝑒𝑟𝑓 √ , (30)
𝑒𝑟𝑓 (𝜂) 2 𝑑𝐺 𝑡
A quasi 1D computational domain with only one grid cell in the direction of translational invariance
is considered. A very thin vapor film is inserted in the computational domain in the initial time near the
hot wall. The liquid-vapor thermo physical properties for water at saturation pressure 1.0 Mpa are cho-
𝜌
sen which correspond to the moderate density ratio of ( 𝜌𝐿 ≈ 170). In order to ensure that the coefficient
𝐺
of mass flux rate in the energy equation is constant in the CFD model during the phase change process,
liquid and vapor phases specific heat are assumed equal (𝐶𝑝𝐺 = 𝐶𝑝𝐿 (𝑃𝑠𝑎𝑡 )). A schematic of Stefan
problem is illustrated in Fig. 8. A no slip boundary condition is employed for the velocity boundary
condition at the hot wall. Temperature of the super-heated wall is 10K higher than saturation temper-
ature.
The integrated simulation error is estimated as the film thickness error (|𝛿𝑠𝑖𝑚 − 𝛿𝑎𝑛 |) summed over
time steps 𝑖, weighted by time step (Δ𝑡):
∑
𝐸= |𝛿𝑠𝑖𝑚 − 𝛿𝑎𝑛 |Δ𝑡. (32)
𝑖
14 SAMKHANIANI AND ANSARI
FIGURE 8 Schematic of Stefan problem, boundary conditions, 𝑃𝑠𝑎𝑡 = 1𝑀𝑝𝑎, 𝑇𝑠𝑎𝑡 = 453.03𝐾, Δ𝑇sup = 10𝐾
FIGURE 9 Interface position in Stefan problem, compare exact solution with present numerical simulations (Lee
and Tanasawa), 𝑃𝑠𝑎𝑡 = 1𝑀𝑝𝑎 and 𝑟𝑒 = 105 , 𝛾 = 1.0
Simulation is conducted on three different grids with 128, 256, and 512 cells. The final interface
position is at 𝑡 = 50(𝑠) and the integrated errors are display in Table 3. It shows the numerical error in
the Tanasawa model is smaller than the Lee model on coarse grids. Furthermore, the numerical error
becomes smaller on finer grids in the Lee model, but remains constant in the Tanasawa model.
𝑇 −𝑇
The interface position (𝛿) and dimensionless temperature (𝜃 = 𝑇 −𝑇𝑤 ) are displayed in Figs. 9 and
𝑠𝑎𝑡 𝑤
10, respectively. It can be seen there is excellent agreement between CFD results and the exact solution.
SAMKHANIANI AND ANSARI 15
FIGURE 10 Dimensionless temperature in Stefan problem, compare exact solution with present numerical sim-
ulations (Lee and Tanasawa), 𝑃𝑠𝑎𝑡 = 1𝑀𝑝𝑎 and 𝑟𝑒 = 105 , 𝛾 = 1.0
FIGURE 11 Hydrodynamic transition in bubble release pattern observed by Reiman and Grigull63
FIGURE 12 Two-dimensional film boiling, dash line is computational domain, 𝑃𝑠𝑎𝑡 = 21.9𝑀𝑝𝑎
For numerical simulation of two-dimensional film boiling, two scenarios have been reported in the
literature:
1. Quasi-periodic
Bubble pinch off occurs individually in the node16,64 or happens sequentially between node and
anti-node23,64 and creates a quasi-periodic pattern.
2. Bubble column
The vapor bubble is connected to the vapor film with a thin neck and creates a mushroom
shape.16,17,22,65 Some researchers have noted that 2D numerical simulations cannot be expected to
give an accurate representation of the 3D physics of film boiling. Furthermore, the two-dimensional
simulation only considers the components of the interface curvature lying in the computational plane,
whereas the out-of-plane components are neglected. In other word, surface tension driving hydrody-
namic instabilities usually cannot be captured in a realistic way17 and Rayleigh–Plateau instability
which plays a role in bubble pinch off may never show up in the 2D model.17,22
The sharp interface method was conducted in all previous studies on the numerical simulation of
film boiling,16,17,22,23,64,66 hence, the present numerical study using the diffuse interface method is
compared with them. Here, the film boiling problems are set similar to Tomar23 and Guo64 works
which are for saturated water at near critical pressure𝑃𝑠𝑎𝑡 = 21.9𝑀𝑝𝑎. They simulated film boiling
using the couple of level set and VOF method (CLSVOF23 and VOSET64 ) with a sharp mass transfer
model. Additionally, Guo64 applied a PLIC method for interface reconstruction to gain more accuracy.
A schematic of two-dimensional film boiling is illustrated in Fig. 12. The super-heated wall temperature
is set 10k and 30K above the saturation temperature. The wall boundary condition is
𝜕𝑃𝑟𝑔ℎ 𝜕𝛼𝐿
⃖⃖⃗ = (0, 0),
at 𝑦 = 0 ∶ 𝑈 𝑇 = 𝑇𝑠𝑎𝑡 + Δ𝑇sup , = = 0. (35)
𝜕𝑦 𝜕𝑦
SAMKHANIANI AND ANSARI 17
𝜕𝑇 𝜕𝛼 ⃖⃖⃗
𝜕𝑈
at 𝑦 = 𝐻 ∶ 𝑃𝑟𝑔ℎ = 𝑃𝑠𝑎𝑡 , = 𝐿 = = 0. (36)
𝜕𝑦 𝜕𝑦 𝜕𝑦
At the beginning of simulation, a vapor film is inserted in the computational domain. Vapor film
thickness is given by66
( )
𝜋𝑁𝑥
𝑦 = 𝑦𝑐 + 𝜀𝐶𝑜𝑠 , (37)
𝑊
where 𝑦𝐶 , 𝜀 and 𝑁 are unperturbed film thickness, perturbation amplitude and perturbation wave,
respectively. The following values are chosen for the present simulation:
4.0 1 /
𝑦𝐶 = 𝜆2𝑑 , 𝜀 = 𝜆2𝑑 , 𝑁 = 1, 𝑊 = 𝜆2𝑑 2, (38)
128 128
where 𝜆2𝑑 is the most unstable Taylor in viscid wavelength for two-dimensional horizontal flat surfaces
and is defined as
√
3𝜎
𝜆2𝑑 = 2𝜋 . (39)
𝑔(𝜌𝐿 − 𝜌𝐺 )
Nusselt number is computed as the dimensionless heat flux through the wall and is given by
𝑊( )
𝜆 𝜕𝑇 |
∫ 𝑇𝑤 −𝑇𝑠𝑎𝑡
× |
𝜕𝑦 |𝑦=0
𝑑𝑥
0
𝑁𝑢2𝑑 = , (40)
𝑊
where 𝑊 is the computational domain width which is 𝜆2𝑑 ∕2. The computational domain height is 𝜆2𝑑
in the present simulations.
The grid independence test is conducted with different grid sizes using the Tanasawa mass transfer
model. Figure 13 depicts the interface as the first bubble is going to leave the vapor film. The bubble
detachment time from the thin vapor neck is dependent on grid size, because the interface breaks
whenever the thickness of neck falls below the specific resolved grids. Therefore, in the coarse grid
(120 × 240) the bubble leaves the vapor film sooner. There is no significant change in bubble shape
between the two last finer grids, so a grid with 240 × 480 cells is chosen for the film boiling simulations.
Figure 14 shows the quasi-periodic release of bubble from node and anti-node. The temperature of
the solid wall is set at 10K above saturation temperature. With the given initial interface, a bubble
is formed at the node (center). Then, the bubble leaves the film under the effect of buoyancy. After
detachment, the vapor neck returns to the wall due to surface tension and moves to the anti-node lead
bubble growth on the outer side. The present numerical result is in good agreement with the numerical
results of Refs. 64 and 23 in predicting the bubble release pattern, bubble shape, and first detachment
time of bubble. The space-averaged Nusselt number over time is shown in Fig. 15. The Nusselt number
depends strongly on the film thickness. Heat flux is greater where the vapor film is thin and less when
the film is the thick. Therefore, average heat flux and Nusselt number increase when the vapor rushes
to fill the bubble and the remaining film becomes thin. To the contrary, after detachment, the vapor
returns to the superheated wall and increased film thickness leads to a smaller Nusselt number. The
Nusselt number predicted by Berenson’s correlation is 3.69. The computed time and space averaged
Nusselt number is 2.57 and 2.64 for the Tanasawa and Lee models, respectively. The difference is
about 28% and 30%. Guo et al.64 and Tomar et al.23 obtained average Nusselt numbers of 4.67 and
18 SAMKHANIANI AND ANSARI
F I G U R E 1 3 Bubble shape at five grids before the first detachment at t = 0.32 s, the grids 210 × 420 and 240 ×
480 have almost the same shape, 𝑃𝑠𝑎𝑡 = 21.9𝑀𝑝𝑎, Δ𝑇sup = 10𝐾
5.37 for a similar problem, corresponding to 27% and 45% difference, respectively. This shows the
present numerical errors are in the range of previous numerical simulations.
The second series of simulation were performed with a wall superheat of 30K above saturation
temperature. Figure 16 shows the interface evolution at different time instants. For this case, Lee’s mass
transfer model still predicts the quasi periodic release of the bubble, but Tanasawa’s mass transfer model
predicts a tall and stable vapor column which is similar with numerical studies of Refs. 23 and 64. The
difference may exist due to how these models calculate transferred mass flux rate during the boiling
process. In Lee’s model, phase change does not need any initial interface. The evaporation takes place
wherever in the liquid phase with the temperature above saturation temperature, thus several small
bubbles can also be seen in the CFD result in Fig. 16 using Lee’s model. In contrast, Tanaswa’s model
is similar to the sharp interface model, mass transfer only occurs across the existing interface. The space
SAMKHANIANI AND ANSARI 19
FIGURE 15 Simulation results (Lee and Tanasawa) for space-averaged Nusselt number compare with Berenson
correlation, 𝑃𝑠𝑎𝑡 = 21.9𝑀𝑝𝑎, Δ𝑇sup = 10𝐾
averaged Nusselt number over time for this case is depicted in Fig. 17. It can be found that when vapor
film is stable, the space averaged Nusselt number is almost uniform. Lee’s model predicts a higher
Nusselt number, because after bubble detachment, the vapor rushes back to the wall and increases the
heat convection mode near the heated wall. The computed time and space averaged Nusselt number
is 2.04 and 2.71 for Tanasawa and Lee models, respectively. The Nusselt number from Berenson’s
correlation is 2.81, therefore, the difference is 44% and 26%, respectively.
FIGURE 17 Simulation results (Lee and Tanasawa) for space-averaged Nusselt number compare with Berenson
correlation, 𝑃𝑠𝑎𝑡 = 21.9𝑀𝑝𝑎, Δ𝑇sup = 30𝐾
In this case, vapor condenses to form a liquid film on the top surfaces of an isothermal plate at 𝑇𝑤 .
By assuming a linear temperature profile from 𝑇𝑤 at the lower wall to 𝑇𝑠𝑎𝑡 at the interface, a control
volume analysis leads to an analytical solution for the film thickness 𝛿(𝑡):33
[ ( )−1 ]1∕2
𝑘 1 𝐻𝐿𝐺
𝛿𝑎𝑛 (𝑡) = 2𝑡 𝐿 + . (41)
𝜌𝐿 𝐶𝑝𝐿 2 𝐶𝑝𝐿 Δ𝑇
Relevant thermo physical properties are for saturated water at 𝑃𝑠𝑎𝑡 = 1𝑀𝑝𝑎. Simulation of fluids
with greater liquid to vapor property ratios and phase change enthalpies require finer meshes and
smaller time steps to adequately resolve the increased range of scales and heat and mass flux mag-
nitudes. Thus, the domain represents a section of an infinitely wide condensing film, a quasi 1D com-
putational domain with only one grid cell in the direction of the translational invariance is considered.
The top free stream boundary is set to saturation temperature 𝑇𝑠𝑎𝑡 = 453.03𝐾 and the bottom wall
is set 𝑇𝑤 = 423.03𝐾 which corresponds to a 30K subcooled temperature. All fluid entering through
SAMKHANIANI AND ANSARI 21
FIGURE 19 Film thickness in horizontal film condensation, compare exact solution with present numerical sim-
ulations (Lee and Tanasawa), 𝑃𝑠𝑎𝑡 = 1𝑀𝑝𝑎 and 𝑟𝑐 = 105 , 𝛾 = 1.0
FIGURE 20 Relative error in horizontal film condensation simulations (Lee and Tanasawa), 𝑃𝑠𝑎𝑡 = 1𝑀𝑝𝑎 and
𝑟𝑐 = 105 , 𝛾 = 1.0
the free stream boundary is vapor. In the beginning of simulation, a very thin film is inserted in the
computational domain.
The development of the liquid film thicknesses and relative errors are presented in Figs. 19 and 20,
respectively. The simulation relative error is defined as
𝛿𝑠𝑖𝑚 − 𝛿𝑎𝑛
𝑟𝐸 = . (42)
𝛿𝑎𝑛
Relative errors in predicted film thickness are higher during the beginning of simulations, because
the liquid films are initially under resolved on fixed meshes. Relative error is smoother and higher in
Lee’s model than Tanasawa’s model. Integrated simulation error was defined in Eq. (32) and presented
in Table 4. These results indicate an excellent agreement between both mass transfer models with the
analytical solution.
22 SAMKHANIANI AND ANSARI
FIGURE 21 Laminar film condensation on a vertical plate, boundary conditions and grid
FIGURE 22 Comparison of film thickness given by present numerical simulations (Tanasawa and Lee) and Nus-
selt’s solution
𝑔(𝜌𝐿 − 𝜌𝐺 ) ( 1
)
𝑈 (𝑦) = 𝛿𝑦 − 𝑦2 . (44)
𝜇𝐿 2
FIGURE 23 Comparison velocity distribution ax z = 5 mm given by present numerical simulations (Lee and
Tanasawa) and Nusselt’s solution
cells in a row near the wall or inlet is 0.1 the height of the cells row near free stream or outlet, respec-
tively. The wall is 20 K subcooled. The boundary conditions and mesh is illustrated at Fig. 21. For the
vapor side, total pressure is assigned for pressure and velocity is calculated from the pressure boundary
condition.
Simulation is conducted for 2 seconds to reach steady state condition. Assuming negligible inter-
shear, the wall shear stress can be evaluated as 𝜏𝑤 = 𝜌𝐿 𝑔𝛿 and dimensionless wall distance as
facial √
𝑔𝛿𝑦
𝑦+ = 𝜈𝐿
where 𝑦 is the cell center distance to the nearest wall and 𝜈𝐿 is the liquid kinematic vis-
cosity. In the present simulation 𝑦+ is smaller than 3. This low value of y+ indicates that near wall
fluid and heat transfer physics are well resolved. Figure 22 shows the liquid film thickness at steady
state determined from numerical result using Tanasawa’s and Lee’s models. It is seen that the liquid
film thickness obtained from the present numerical simulation agrees well with the classical Nusselt
analytical prediction.
The velocity profile from the numerical simulation is compared with Nusselt’s solution in Fig. 23.
The velocity increases from zero at the wall to a maximum value away from the wall and then decreases
monotonically to zero in the vapor at infinity. Thus, the velocity profile is continuous from the sub-
cooled liquid at the wall to dry vapor at infinity. The velocity profile given by the Nusselt model is
discontinuous. It can be seen that the velocity obtained by both mass transfer models are the same and
they are smaller than that of the classical Nusselt model given by Eq. (44). Because the classical Nus-
selt model did not take into consideration the effects of inertia forces, convection terms and interfaces
shear stress. In contrast, these three parameters are taken into consideration in the numerical simula-
tions. Therefore, the film liquid induces the vapor flow and at the same time, liquid film gives up some
of its momentum, and therefore its velocity decreases as compared to that given by Eq. (44). A similar
result for the velocity profile is pointed out in numerical results of Liu and Cheng69 who simulated
laminar film condensation with the Lattice Boltzman method.
𝑇 −𝑇
The dimensionless temperature defined as 𝜃 = 𝑇 −𝑇𝑤 is presented in Fig. 24. Nusselt’s solution
𝑠𝑎𝑡 𝑤
assumed a linear variation of temperature within the liquid film and constant saturation temperature
outside of the condensate liquid film. Numerical simulation shows for the present thermophysical prop-
erties and film thickness the linear variation of temperature in the liquid film is not valid.
SAMKHANIANI AND ANSARI 25
FIGURE 24 Temperature profiles at z = 5 mm in film condensation on a vertical flat plate, compare present
numerical simulations (Tanasawa and Lee) and Nusselt’s solution
FIGURE 25 Schematic of bubble condensation in subcooled boiling, initial and boundary conditions
FIGURE 26 Bubble shape at 𝑡 = 1𝑚𝑠 for different grids, 𝑃𝑠𝑎𝑡 = 0.130[𝑀𝑝𝑎] and Δ𝑇𝑠𝑢𝑏 = 25[𝐾]
FIGURE 27 Bubble diameter history, compare between present numerical simulations and experimental data,70
𝐷0 = 1.008𝑚𝑚, 𝑃𝑠𝑎𝑡 = 0.130[𝑀𝑝𝑎] and Δ𝑇𝑠𝑢𝑏 = 25[𝐾]
in Fig. 25 which is similar to Samkhaniani and Ansari.71 The 2D space domain is set as 2𝐷0 × 4𝐷0
where 𝐷0 is the initial diameter of vapor bubble. The bubble is located in the position of (𝐷0 , 𝐷0 ) at
the beginning of the simulation.
A mesh resolution analysis is performed on four different grids using Tanasawa’s model to figure
out the appropriate grid size for simulation of bubble condensation. These grids are 50 × 100, 75 ×
150, 100 × 200, and 150 × 300 which, respectively, represent 25, 37, 50, and 75 cells across the initial
diameter. It can be seen obviously in Fig. 26 that bubble shape becomes stable with meshes of 100 ×
200 and 150 × 300. The variation between the two lateral grids is negligible. Therefore, the bubble is
resolved using 100 × 200 cells in this study.
In order to validate the present numerical simulation, the shape and the bubble diameter history
is compared with published experimental results70 under saturation pressure 0.130 Mpa for water
SAMKHANIANI AND ANSARI 27
FIGURE 28 Bubble shape, comparison between present numerical results (Lee and Tanasawa) and experimental
data,70 𝐷0 = 1.008𝑚𝑚, 𝑃𝑠𝑎𝑡 = 0.130[𝑀𝑝𝑎] and Δ𝑇𝑠𝑢𝑏 = 25[𝐾]
𝜌
corresponding to the severe density ratio of ( 𝜌𝐿 ≈ 1263). The comparison of bubble diameter history
𝐺
is shown in Fig. 27 and the bubble shape in Fig. 28. Bubble diameter in 2D simulation is the diameter
of an equivalent circle. It can be seen that numerical result is in reasonable good agreement with the
experimental results.
4. CONC LU SI ON
In the present study, a new solver called phaseChangeHeatFoam is implemented in the OpenFOAM
cfd package to simulate boiling and condensation processes. For now, two mass transfer models (Lee
and Tanasawa) are implemented in the solver. However, adding a new mass transfer model is straight
forward. The governing equations and programming information are discussed in detail. This solver is
based on the CF-VOF method, but spurious current is reduced by using a filter method for calculation
of the interface curvature. The present numerical simulations are compared with analytical solutions,
previous sharp interface method results and experimental data for different thermo-physical properties.
Our investigation shows the present diffuse interface method is capable of accurate simulation of the
𝜌
phase change process even for a severe ratio of thermophysical properties ( 𝜌𝐿 ≈ 1000). Additionally,
𝑣
the performance of Lee and Tanasawa mass transfer models are investigated here through simulations,
it was found both mass transfer models have similar performance in simple simulations (Stefan and
horizontal film condensation). However, in the simulation of complex phenomena, for instance film
boiling, the difference in the numerical results obtained with Lee and Tanasawa mass transfer models
can be considerable. Our study shows numerical results with Tanasawa mass transfer model are very
similar to the sharp interface model.
28 SAMKHANIANI AND ANSARI
Liquid
𝒌𝒈
𝑻 [𝑲] 𝑷 [𝑴𝒑𝒂] 𝝆[ 𝒎𝟑 ] 𝑯[ 𝒌𝑱
𝒌𝒈
] 𝒌𝑱
𝑪𝒑 [ 𝒌𝒈.𝑲 ] 𝝁[𝑷 𝒂.𝒔] 𝒘
𝒌[ 𝒎.𝑲 ] 𝝈[ 𝑵
𝒎
]
380.26 0.13 953.13 449.19 4.2244 2.620E-04 0.68106 0.05753
453.03 1 887.13 762.52 4.4045 1.5024E-4 0.67337 0.042217
646 21.9 402.4 1963.5 218 4.67E-05 0.545 0.00007
Vapor
𝒌𝒈
𝑻 [𝑲] 𝑷 [𝑴𝒑𝒂] 𝝆[ 𝒎𝟑 ] 𝑯[ 𝒌𝑱
𝒌𝒈
] 𝒌𝑱
𝑪𝒑 [ 𝒌𝒈.𝑲 ] 𝝁[𝑷 𝒂.𝒔] 𝒘
𝒌[ 𝒎.𝑲 ]
380.26 0.13 0.7545 2686.6 2.1107 1.251E-05 0.02590
453.03 1 5.1450 2777.1 2.7114 1.502E-05 0.03643
646 21.9 242.7 2240 352 3.23E-06 0.538
REFE RENCES
1. Gingold RA, Monaghan JJ. Smoothed particle hydrodynamics: theory and application to non-spherical stars. Mon
Not R Astron Soc. 1977;181:375–389.
2. Unverdi SO, Tryggvason G. A front-tracking method for viscous, incompressible, multi-fluid flows. J Comput Phys.
1992;100:25–37.
3. Sussman M, Smereka P, Osher S. A level set approach for computing solutions to incompressible two-phase flow.
J Comput Phys. 1995;114:146–159.
4. Hirt CW, Nichols BD. Volume of fluid (VOF) method for the dynamics of free boundaries. J Comput Phys.
1981;39:201–225.
5. Worner M. Numerical modeling of multiphase flow in microfluidics and micro process engineering: a review of
methods and applications. Microfluid NanoFluid. 2012;12:841–886.
6. Sussman M, Puckett EG. A coupled level set and volume-of-fluid method for computing 3D and axisymmetric
incompressible two-phase flows. J Comput Phys. 2000;162:301–337.
7. Sussman M. A second order coupled level set and volume-of-fluid method for computing growth and collapse of
vapor bubbles. J Comput Phys. 2003;187:110–136.
8. Van der Pijl S, Segal A, Vuik C, Wesseling P. A mass conserving level set method for modelling of multi-phase
flows. Int J Numer Methods Fluids. 2005;47:339–361.
9. Lv X, Zou Q, Zhao Y, Reeve D. A novel coupled level set and volume of fluid method for sharp interface capturing
on 3D tetrahedral grids. J Comput Phys. 2010;229:2573–2604.
10. Sun D, Tao W. A coupled volume-of-fluid and level set (VOSET) method for computing incompressible two-phase
flows. Int J Heat Mass Transfer. 2010;53:645–655.
11. Youngs DL. Time-dependent multi-material flow with large fluid distortion. Num Methods Fluid Dyn. 1982;24:273–
285.
12. Renardy Y, Renardy M. PROST: a parabolic reconstruction of surface tension for the volume-of-fluid method.
J Comput Phys. 2002;183:400–421.
13. Ubbink O, Issa R. A method for capturing sharp fluid interfaces on arbitrary meshes. J Comput Phys. 1999;153:26–
50.
14. Xiao F, Honma Y, Kono T. A simple algebraic interface capturing scheme using hyperbolic tangent function. Int
J Numer Methods Fluids. 2005;48:1023–1040.
15. Lyklema J. Fundamentals of Interfaces and Colloid Science, Vol. I: Fundamentals. Academic Press; 1991.
16. Welch SWJ, Wilson J. A volume of fluid based method for fluid flows with phase change. J Comput Phys.
2000;160:662–682.
SAMKHANIANI AND ANSARI 29
17. Hardt S, Wondra F. Evaporation model for interfacial flows based on a continuum-field representation of the source
terms. J Comput Phys. 2008;227:5871–5895.
18. Schlottke J, Weigand B. Direct numerical simulation of evaporating droplets. J Comput Phys. 2008;227 5215–5237.
19. Son G, Dhir V. Numerical simulation of film boiling near critical pressures with a level set method. J Heat Transfer.
1998;120:183–192.
20. Son G, Dhir V, Ramanujapu N. Dynamics and heat transfer associated with a single bubble during nucleate boiling
on a horizontal surface. J Heat Transfer. 1999;121:623–631.
21. Tanguy S, Ménard T, Berlemont A. A level set method for vaporizing two-phase flows. J Comput Phys.
2007;221:837–853.
22. Gibou FDR, Chen L, Nguyen D, Banerjee S. A level set based sharp interface method for the multiphase incom-
pressible Navier–Stokes equations with phase change. Comput Phys. 2007;222:536–555.
23. Tomar G, Biswas G, Sharma A, Agrawal A. Numerical simulation of bubble growth in film boiling using a coupled
level-set and volume-of-fluid method. Phys Fluids. 2005;17.
24. Kunkelmann C, Stephan P. Numerical simulation of the transient heat transfer during nucleate boiling of refrigerant
HFE-7100. Int J Refrig. 2010;33:1221–1228.
25. Weller HG, TaboraI G, Jasak H, Fureby C. A tensorial approach to computational continuum mechanics using
object-oriented techniques. Comput Phys. 1998;12:620–632.
26. Deshpande SS, Anumolu L, Trujillo MF. Evaluating the performance of the two-phase flow solver interFoam.
Comput Sci Discovery. 2012;5:014016.
27. Berberović E, Hinsberg NPv, Jakirlić S, Roisman IV, Tropea C. Drop impact onto a liquid layer of finite thickness:
dynamics of the cavity evolution. Am Phys Soc. 2009;79:036306 (15).
28. Klostermann J, Schaake K, Schwarze R. Numerical simulation of a single rising bubble by VOF with surface com-
pression. Int J Numer Methods Fluids. 2013;71:960–982.
29. Ubbink O. Numerical Prediction of Two Fluid Systems with Sharp Interfaces. London: University of London; 1997.
30. Kunkelmann C, Stephan P. CFD simulation of boiling flows using the volume-of-fluid method within OpenFOAM.
Num Heat Transfer 2009;56:631–646.
31. Zeng Q, Cai J, Yin H, Yang X, Watanabe T. Numerical simulation of single bubble condensation in subcooled flow
using OpenFOAM. Progress Nucl Energy. 2015;83:336–346.
32. Doro EO. Computational modeling of falling liquid film free surface evaporation. 2012.
33. Rattner AS, Garimella S. Simple mechanistically consistent formulation for volume-of-fluid based computations of
condensing flows. J Heat Transfer. 2014;136:071501.
34. Bahreini M, Ramiar A, Ranjbar AA. Numerical simulation of bubble behavior in subcooled flow boiling under
velocity and temperature gradient. Nucl Eng Des. 2015;293:238–248.
35. Schrage RW. A Theoretical Study of Interphase Mass Transfer. Columbia University Press; 1953.
36. Knudsen M. The Kinetic Theory of Gases: Some Modern Aspects. Metheun & Company; 1950.
37. Marek R, Straub J. Analysis of the evaporation coefficient and the condensation coefficient of water. Int J Heat Mass
Transfer. 2001;44:39–53.
38. Tanasawa I. Advances in condensation heat transfer. Adv Heat Transfer. 1991;21:55–139.
39. Lee WH. A pressure iteration scheme for two-phase flow modeling. Multiph Transp Fundam React Saf Appl.
1980;1:407–431.
40. Sun D-L, Xu J-L, Wang L. Development of a vapor–liquid phase change model for volume-of-fluid method in
FLUENT. Int. Commun Heat Mass Transfer. 2012;39:1101–1106.
41. Wu H, Peng X, Ye P, Gong YE. Simulation of refrigerant flow boiling in serpentine tubes. Int J Heat Mass Transfer.
2007;50:1186–1195.
42. De Schepper SC, Heynderickx GJ, Marin GB. Modeling the evaporation of a hydrocarbon feedstock in the convec-
tion section of a steam cracker. Comput Chem Eng. 2009;33:122–132.
30 SAMKHANIANI AND ANSARI
43. Alizadehdakhel A, Rahimi M, Alsairafi AA. CFD modeling of flow and heat transfer in a thermosyphon. Int Commun
Heat Mass Transfer. 2010;37:312–318.
44. Yang Z, Peng X, Ye P. Numerical and experimental investigation of two phase flow during boiling in a coiled tube.
Int J Heat Mass Transfer. 2008;51:1003–1016.
45. Goodson K, Rogacs A, David M, Fang C. Volume of fluid simulation of boiling two-phase flow in a vapor-venting
microchannel. Frontiers Heat Mass Transfer. 2010;1.
46. Lee H, Kharangate CR, Mascarenhas N, Park I, Mudawar I. Experimental and computational investigation of vertical
downflow condensation. Int J Heat Mass Transfer. 2015;85:865–879.
47. Albadawi A, Donoghue D, Robinson A, Murray D, Delauré Y. Influence of surface tension implementation in
volume of fluid and coupled volume of fluid with level set methods for bubble growth and detachment. Int J Multiph
Flow. 2013;53:11–28.
48. Samkhaniani N, Gharehbaghi A, Ahmadi Z. Numerical simulation of reaction injection molding with polyurethane
foam. J Cell Plast. 2013;49:405–421.
49. Marschall H, Hinterberger K, Schuler C, Habla F, Hinrichsen O. Numerical simulation of species transfer across
fluid interfaces in free-surface flows using OpenFOAM. Chem Eng Sci. 2012;78:111–127.
50. Brackbill JU, Kothe DB, Zemach C. A continuum method for modeling surface tension. J Comput Phys.
1992;100:335–354.
51. Scardovelli R, Zaleski S. Direct numerical simulation of free-surface and interfacial flow. Ann Rev Fluid Mech.
1999;31:567–603.
52. Lafaurie B, Nardone C, Scardovelli R, Zaleski S, Zanetti G. Modelling merging and fragmentation in multiphase
flows with SURFER. J Comput Phys. 1994;113:134–147.
53. Hoang DA, van Steijn V, Portela LM, Kreutzer MT, Kleijn CR. Benchmark numerical simulations of segmented
two-phase flows in microchannels using the volume of fluid method. Comput Fluids. 2013;86:28–36.
54. Zalesak ST. Fully multidimensional flux-corrected transport algorithms for fluids. J Comput Phys. 1979;31:335–
362.
55. Issa RI. Solution of the implicitly discretised fluid flow equations by operator-splitting. J Comput Phys. 1986;62:40–
65.
56. Rhie C, Chow W. Numerical study of the turbulent flow past an airfoil with trailing edge separation. AIAA J.
1983;21:1525–1532.
57. Van Leer B. Towards the ultimate conservative difference scheme. II. Monotonicity and conservation combined in
a second-order scheme. J Comput Phys. 1974;14:361–370.
58. Weller H. A new approach to VOF-based interface capturing methods for incompressible and compressible flow.
OpenCFD Ltd, Report TR/HGW/04. 2008.
59. Jasak H. Error Analysis and Estimation for the Finite Volume Method with Applications to Fluid Flows: University
of London; 1996.
60. Sun D-L, Xu J-L, Wang L. Development of a vapor–liquid phase change model for volume-of-fluid method in
FLUENT. International Communications in Heat and Mass Transfer 2012;39:1101–1106.
61. Tsujimoto K, Kambayashi Y, Shakouchi T, Ando T. Numerical simulation of gas-liquid two-phase flow with phase
change using Cahn-Hilliard equation. Turbulence Heat Mass Transfer 2009;6.
62. Berenson P. Film-boiling heat transfer from a horizontal surface. J Heat Transfer. 1961;83:351–356.
63. Reimann M, Grigull U. Warmeubergang bei freier konvektion und flimsieden im kritischen gebiet von wasser und
kohlendioxid. Warmeund Stof-fubertragung. 1975;8:229–239.
64. Guo DZ, Sun DL, Li ZY, Tao WQ. Phase change heat transfer simulation for boiling bubbles arising from a vapor
film by the VOSET method. Num Heat Transfer. 2011;59:857–881.
65. Juric D, Tryggvason Gt. Computations of boiling flows. Int J Multiphase Flow. 1998;24:387–410.
66. Esmaeeli A, Tryggvason G. Computations of film boiling. Part I: numerical method. Int J Heat Mass Transfer.
2004;47
SAMKHANIANI AND ANSARI 31
67. Ghiaasiaan SM. Two-Phase Flow, Boiling and Condensation in Conventional and Miniature System. Cambridge
University Press; 2008.
68. Nusselt W. The surface condensation of water vapour. VDI Z. 1916;60:541–546.
69. Liu X, Cheng P. Lattice Boltzmann simulation of steady laminar film condensation on a vertical hydrophilic sub-
cooled flat plate. Int J Heat Mass Transfer. 2013;62:507–514.
70. Kamei S, Hirata M. Condensing phenomena of a single vapor bubble into subcooled water. Exp Heat Transfer.
1990;3:173–182.
71. Samkhaniani N, Ansari M. Numerical simulation of bubble condensation using CF-VOF. Progress in Nuclear
Energy. 2016;89:120–131.
How to cite this article: Samkhaniani N, Ansari MR. The evaluation of the diffuse inter-
face method for phase change simulations using OpenFOAM. Heat Transfer–Asian Research.
2017;00:1–31. doi:10.1002/htj.21268