Euler-Euler Approach
In this approach, the multiphase VOF method is applied to simulate the transient three-dimensional air/water flow in a water cylinder, and argon gas/steel/slag/air four-phase fluid flow in an LMF.
The VOF formulation relies on the fact that two or more fluids (or phases) are not interpenetrating. For each additional phase that is added to the model, a variable is introduced, which represent the volume fraction of the phase in the computational cell. In each control volume, the volume fractions of all phases sum to unity. The tracking of the interfaces among phases is accomplished by solving the continuity equation for the volume fraction of phases. The continuity equation for the qth phase is described in the following form:[7,25]
$$ \frac{1}{{\rho_{q} }}\left[ {\frac{\partial }{\partial t}\left( {\alpha_{q} \rho_{q} } \right) + \nabla \cdot \left( {\alpha_{q} \rho_{q} \overrightarrow {{u_{q} }} } \right)} \right] = 0, $$
(1)
where the volume fraction α
q
is constrained by \( \mathop \sum\nolimits_{q = 1}^{n} \alpha_{q} = 1 \).
The momentum equation is solved throughout the domain, and the resulting velocity field is shared among the phases. The momentum equation expressed as Eq. [2] is dependent on the volume fractions of all phases.
$$ \frac{\partial }{\partial t}\left( {\rho \vec{v}} \right) + \nabla \cdot \left( {\rho \vec{v}\vec{v}} \right) = - \nabla p + \nabla \cdot \left[ {\mu \left( {\nabla \vec{v} + \nabla \vec{v}^{T} } \right)} \right] + \rho \vec{g}, $$
(2)
where ρ and μ are the density and viscosity, respectively, of the mixture depending on the volume fraction α
q
of each phase. \( \vec{v} \) is the underlying velocity field and p is the local pressure.
Euler-Lagrange Approach
In this approach, the VOF model is used for describing continuous phases (slag, steel, and air phases) and tracking the interfaces among phases. The interfaces are captured by the geo-reconstruct scheme, and the area of the slag-steel interface is computed and recorded every time step during the simulation. The blowing gas is treated as a discrete phase by injecting a stream of gas bubbles into the continuous phase.
The trajectory of each bubble is calculated in each time step according to force balance of the drag force, buoyancy force, virtual mass force, and the pressure gradient force.
$$ \frac{{d\vec{u}_{p} }}{dt} = F_{\text{D}} \left( {\vec{u} - \vec{u}_{p} } \right) + \frac{{\vec{g}\left( {\rho_{p} - \rho } \right)}}{{\rho_{p} }} + \vec{F}_{\text{vm}} + \vec{F}_{P} .$$
(3)
The four terms on the right represent the contributions of drag force, gravity, virtual mass, and pressure gradient force to particle acceleration. FD is written as
$$ F_{\text{D}} = \frac{18\mu }{{\rho_{p} d_{p}^{2} }}\frac{{C_{D} Re}}{24}. $$
(4)
The virtual mass force, which would accelerate the fluid surrounding the bubble, is expressed as
$$ \vec{F}_{\text{vm}} = C_{\text{vm}} \frac{\rho }{{\rho_{p} }}\left( {\vec{u}_{p} \nabla \vec{u} - \frac{{d\vec{u}_{p} }}{dt}} \right). $$
(5)
The force that comes from the pressure gradient in the fluid is described as
$$ \vec{F}_{P} = \frac{\rho }{{\rho_{p} }}\vec{u}_{p} \nabla \vec{u}. $$
(6)
Here, \( \vec{u} \) is the fluid phase velocity; \( \vec{u}_{p} \) is bubble velocity; ρ is the fluid density; ρ
p
is the density of the bubble; \( \mu \) is the viscosity of the fluid, and d
p
is the bubble diameter. Re is the relative Reynolds number. Cvm is the virtual mass factor with a default value of 0.5. The drag coefficient, CD, is calculated as a function of bubble shape by using Eotvos number:[17]
$$ C_{\text{D}} = \frac{2}{3}\left( {\frac{{E_{0} }}{3}} \right)^{0.5} $$
(7)
$$ E_{0} = \frac{{g\left( {\rho - \rho_{p} } \right)d_{p}^{2} }}{\sigma }, $$
(8)
where Eotvos number, E0, is a dimensionless number describing the shape of the bubble. σ is the surface tension between the steel and the gas.
Turbulent dispersion creates a random addition to the liquid velocity for the drag force in Eq. [3]. It results in a wider bubble plume. Discrete random walk model is applied to account for the effects of turbulent dispersion on bubbles. Two-way turbulence coupling is implemented between the discrete and the continuous phases to facilitate momentum transfer between the bubbles and the continuous phases.
The collision, coalescence, and breakup phenomena of bubbles are included in this model. The algorithm of O’Rourke[26] is used to describe the coalescence process of the bubbles. Bubbles are considered to coalescence if they collide head-on, or to bounce if the collision is more oblique. The probability of coalescence can be related to the offset of the collector bubble center and the trajectory of the smaller bubble. And the critical offset is a function of the collisional Weber number and the relative radii of the collector (r1) and the smaller bubble (r2):
$$ b_{\text{crit}} = \left( {r_{1} + r_{2} } \right)\sqrt {\hbox{min} \left( {1.0,\frac{2.4f}{{W_{e} }}} \right)}, $$
(9)
where f is a function of \( r_{1} /r_{2} \), defined as
$$ f\left( {\frac{{r_{1} }}{{r_{2} }}} \right) = \left( {\frac{{r_{1} }}{{r_{2} }}} \right)^{3} - 2.4\left( {\frac{{r_{1} }}{{r_{2} }}} \right)^{2} + 2.7\left( {\frac{{r_{1} }}{{r_{2} }}} \right). $$
(10)
The bubble breakup model proposed by Cloete et al.[17] is applied in the present model, which assumes that when the bubble diameter is over a critical diameter of 40 mm, two bubbles with equivalent mass are generated from one mother bubble. A more sophisticated bubble expansion and breakup model[13] can be used in a future study. The bubbles will disappear and join to the air to be a continuum after arriving to the liquid surface; thus, the present model removes the bubbles when they arrive to a position where the air volume fraction is above 0.5.
Turbulence Model
The realizable k-ε turbulence model is chosen to account for multiphase turbulence flow in both of the above-described CFD models. The standard wall functions are used as near-wall treatments for wall-bounded turbulent flows.[27]
The following transport equations for k and ε are solved:
$$ \frac{\partial }{{\partial {\text{t}}}}\left( {\rho k} \right) + \frac{\partial }{{\partial x_{j} }}\left( {\rho ku_{j} } \right) = \frac{\partial }{{\partial x_{j} }}\left[ {\left( {\mu + \frac{{\mu_{t} }}{{\sigma_{k} }}} \right)\frac{\partial k}{{\partial x_{j} }}} \right] + G_{k} + G_{b} - \rho \varepsilon $$
(11)
$$ \frac{\partial }{\partial t}\left( {\rho \varepsilon } \right) + \frac{\partial }{{\partial x_{j} }}\left( {\rho \varepsilon u_{j} } \right) = \frac{\partial }{{\partial x_{j} }}\left[ {\left( {\mu + \frac{{\mu_{t} }}{{\sigma_{\varepsilon } }}} \right)\frac{\partial \varepsilon }{{\partial x_{j} }}} \right] + \rho C_{1} S\varepsilon - \rho C_{2} \frac{{\varepsilon^{2} }}{{k + \sqrt {v\varepsilon } }} + C_{1\varepsilon } \frac{\varepsilon }{k}C_{3\varepsilon } G_{b}, $$
(12)
where
$$ C_{1} = \hbox{max} \left[ {0.43,\frac{\eta }{\eta + 5}} \right], \,\,\eta = S\frac{k}{\varepsilon }, \,\,S = \sqrt {2S_{ij} S_{ij} } ,\, {\text{and }}\,S_{ij} = \frac{1}{2}\left( {\frac{{\partial u_{j} }}{{\partial x_{i} }} + \frac{{\partial u_{i} }}{{\partial x_{j} }}} \right). $$
In these equations, Gk represents the generation of turbulence kinetic energy due to the mean velocity gradient, calculated as
$$ G_{\text{k}} = - \rho \overline{{u^{\prime}_{i} u^{\prime}_{j} }} \frac{{\partial u_{j} }}{{\partial x_{i} }}. $$
(13)
Gb is the generation of turbulence kinetic energy due to buoyancy, calculated as
$$ G_{b} = \beta g_{i} \frac{{\mu_{t} }}{{Pr_{t} }}\frac{\partial T}{{\partial x_{i} }}, $$
(14)
where Pr
t
is the turbulent Prandtl number for energy, which is 0.85 for the realizable k-ε model. g
i
is the component of the gravitational vector in the th direction. The coefficient of thermal expansion, β, is defined as
$$ \beta = - \frac{1}{\rho }\left( {\frac{\partial \rho }{\partial T}} \right)_{P}. $$
(15)
The eddy viscosity, μ
t
, is computed from
$$ \mu_{t} = \rho C_{\mu } \frac{{k^{2} }}{\varepsilon }. $$
(16)
The degree to which ɛ is affected by the buoyancy is determined by the constant C3ɛ, which is calculated according to the following relation:
$$ C_{3\varepsilon } = { \tanh }\left| {\frac{v}{u}} \right|, $$
(17)
where v is the component of the flow velocity parallel to the gravitational vector and u is the component of the flow velocity perpendicular to the gravitational vector.
The model constants are \( C_{1\varepsilon } = 1.44,\,\,C_{2} = 1.9,\,\, \sigma_{k} = 1.0,\,\, \sigma_{\varepsilon } = 1.2 \). \( C_{\mu } \) is a function of the turbulence fields, as described in Reference 27.
The mass transfer coefficient in steel, km, can be calculated through the Kolmogorov theory of isotropic turbulence as follows:
$$ k_{\text{m}} = cD_{\text{m}}^{0.5} \left( {\frac{{\varepsilon_{\text{l}} }}{{\nu_{\text{l}} }}} \right)^{0.25}, $$
(18)
where c is a constant that is 0.4 for this study.[28] Dm represents the diffusion coefficient of the species in liquid steel as described by Lou et al.[28] \( \varepsilon_{\text{l}} \) is the turbulent energy dissipation rate and \( \nu_{\text{l}} \) is the kinematic viscosity.
Initial and Boundary Conditions
The meshed geometry of the LMF is shown in Figure 1. Initially, the slag layer rests on the top of the steel bath, and no argon blows through the plugs. Non-slip conditions with the standard wall function are employed at the bottom and the side walls. The pressure outlet boundary condition is used at the top surface of the ladle, where the argon gas are allowed to escape. The velocity inlet is used for the gas flow at the bottom plugs. The inlet velocity, ub, is calculated in terms of the gas flow rate, Q.[27]
For the Lagrangian DPM, the number of injected bubbles per second (n) is calculated by using the gas flow rate and the injected bubble size:
$$ n = \frac{6Q}{{\pi d_{p,0}^{3} }}. $$
(19)
The injected bubble size, dp,0, is determined by the following equation:[13]
$$ d_{p,0} = 0.091\left( {\frac{\sigma }{\rho }} \right)^{0.5} u_{b}^{0.44}. $$
(20)
Table I shows the number of the injected bubbles per second under different gas flow rates and the injected bubble size in the DPM simulation studies. It can be seen that the number of bubbles included in the simulation studies is quite large. The computational cost of tracking all of the particles individually is prohibitive. Therefore, in ANSYS Fluent, the DPM model tracks a number of ‘parcels’, and each parcel is a representative of a fraction of the total mass flow released in a time step. The trajectory of each parcel is determined by tracking a single representative particle in the parcel.[16]
Table I The Number of the Injected Bubbles Per Second Under Different Gas Flow Rates and Injected Bubble Size Used in the Simulation Studies
Numerical Procedure
The computation is carried out by using a transient pressure-based solver. The PISO scheme available in ANSYS Fluent is used for the pressure-velocity coupling and the volume fraction equation is solved using an explicit geo-reconstruct scheme.[16] The pressure staggering option (PRESTO!) scheme is selected for the pressure equation. A mesh sensitivity analysis for the VOF model has been conducted by using three different meshes, which correspond with the number of control volume elements of 239,000, 370,000, and 475,000. It was found that the meshes with 370,000 and 475,000 elements produced quite similar simulation results (such as the open slag eyes, slag-steel interface, and velocity distribution inside the ladle). Thus, the mesh with 370,000 elements is suitable for the current numerical simulation studies.
Model Validation
The prediction accuracy of the Euler-Euler and Euler-Lagrange approaches was investigated by comparing the simulated flow characteristics with the experimental results of a water model measured by Sheng et al.[29] The detailed study of the LMF operation simulation was conducted in an industry-scale ladle. The argon gas is supplied through two off-centered plugs at the bottom of the ladle. The ladle dimensions and thermo-physical properties of the steel and the slag are shown in Table II.
Table II The Dimensions of the Ladle and Other Parameters Employed in the Models