Download as pdf or txt
Download as pdf or txt
You are on page 1of 18

metals

Article
Effects of Impeller Rotational Speed and Immersion Depth on
Flow Pattern, Mixing and Interface Characteristics for Kanbara
Reactors Using VOF-SMM Simulations
Qiang Li 1,2, * , Suwei Ma 2 , Xiaoyang Shen 3 , Mingming Li 1,2 and Zongshu Zou 1,2

1 Key Laboratory for Ecological Metallurgy of Multimetallic Mineral (Ministry of Education),


Northeastern University, Shenyang 110819, China; limm@smm.neu.edu.cn (M.L.);
zouzs@mail.neu.edu.cn (Z.Z.)
2 School of Metallurgy, Northeastern University, Shenyang 110819, China; masw0908@gmail.com
3 PERIC Special Gases Co., Ltd., Handan 056107, China; shenxy1006@gmail.com
* Correspondence: liq@smm.neu.edu.cn; Tel.: +86-24-8368-7724

Abstract: The Kanbara Reactor (KR) is a primary desulfurization technology in the hot metal pretreat-
ment refining process that is widely employed in the modern steelmaking industry. The operating
parameters of KR impeller immersion depth (IID) and rotation speed (IRS) have a crucial impact on
the process performance and the desulfurization effect. Still, their influences have not been fully un-
derstood. This study systematically investigated the effects of IID and IRS on the flow pattern, mixing
behavior, vortex core depth, and free surface characteristics for KR processes based on a 3D Volume
of Fluid (VOF) model coupled with the sliding mesh method (SMM). The model was validated via
 scale-down water model experiments and then applied to the KR process, and simulations found

that IID and IRS have different impacts on the flow pattern. Specifically, the discharge flow location
Citation: Li, Q.; Ma, S.; Shen, X.; moves downward with IID increasing, but the discharge strength and mean velocity hardly changes.
Li, M.; Zou, Z. Effects of Impeller Comparatively, the rise of IRS significantly increases the mean velocity, but few changes occur to the
Rotational Speed and Immersion
discharge flow position. Increasing IRS improves bath hydrodynamics, strengthens recirculation,
Depth on Flow Pattern, Mixing and
and efficiently shortens mixing time, but IID has a neglectable effect on these features. The minimum
Interface Characteristics for Kanbara
mixing time is 55 s at a maximum IRS of 260 rpm. Moreover, the vortex core depth and free surface
Reactors Using VOF-SMM
Simulations. Metals 2021, 11, 1596.
velocity visibly increase with the increase of IRS. Comparatively, IID has a limited effect on the flow
https://doi.org/10.3390/met11101596 and mixing behavior but directly impacts the distribution of recirculation regions at the axial direction
and the velocity gradient on the free surface at the radial direction. Furthermore, the correlation
Academic Editor: Diego Celentano equations of these critical parameters as a function of the operating parameters were obtained. The
results from this study may provide references for operating optimizations and industrial practices
Received: 9 September 2021 of KRs.
Accepted: 5 October 2021
Published: 8 October 2021 Keywords: desulfurization; Kanbara Reactor; immersion depth; rotation speed; numerical simulation

Publisher’s Note: MDPI stays neutral


with regard to jurisdictional claims in
published maps and institutional affil- 1. Introduction
iations.
With the increasing demand for high-quality steel, cleanliness is of paramount im-
portance in each stage of steel refining and production. Except for special applications,
the sulfur existing in products is usually a detrimental impurity that adversely affects the
temperature toughness, weldability, and hydrogen-induced cracking resistance during the
Copyright: © 2021 by the authors.
service life of products [1,2]. Thus, it is essential to reduce the sulfur content to as low
Licensee MDPI, Basel, Switzerland.
as possible to satisfy the specification of the product. Desulfurization is usually carried
This article is an open access article
out before the hot metal is mixed into the converter because of favorable thermodynamic
distributed under the terms and
conditions, and this stage is thus called the pretreatment procedure for desulfurization. Ap-
conditions of the Creative Commons
proaches used here can be classified into two categories: mechanically stirred and powder
Attribution (CC BY) license (https://
injection (PI). Among them, the Kanbara Reactor (KR) [3], as a typical mechanically stirred
creativecommons.org/licenses/by/
4.0/).
method (MSM) [4–6], has significant advantages with high efficiency, short treatment time,

Metals 2021, 11, 1596. https://doi.org/10.3390/met11101596 https://www.mdpi.com/journal/metals


Metals 2021, 11, 1596 2 of 18

and lower cost, providing the intensely hydrodynamic conditions through the revolving
motion of a submerged impeller. Practically, the rotational speed and immersion depth of
the impeller, as two vital operating parameters, remarkably impact the entrainment of the
reagent powder, mixing characteristics, and refining efficiency, as well as the service life
of an impeller and the ladle because of higher refractory wear induced by the washing of
revolving, high-temperature metal.
Some studies have been conducted on different aspects of MSM or KRs, including fluid
flows, bath mixing behaviors, reagent dispersion, and impeller design based on scale-down
water models, numerical simulations, and laboratory-scale high-temperature experiments.
For example, Visuri et al. [1] have given a recent review in modeling hot metal desulfur-
ization and summarized the previous contributions and work on fluid flow phenomena,
mixing patterns, and particle/bubble dispersions in KR processes. Yamamoto et al. [7,8]
investigated the flow characteristic of aluminum melt by mechanically stirring and found
that the impeller depths, rotation speeds, and impeller twisting angles impact the mixing
time. Kato et al. [9] evaluated the flow pattern and mass transfer in an aluminum melting
furnace using a numerical simulation. They found that the region with high kinetic energy
was enlarged, and the turbulent kinetic energy at the free surface increased significantly
with impeller rotation speed. This result illustrated that increasing impeller rotation speed
is an effective method to enhance the mass transfer behavior of reactors. Their more re-
cent research [10] found that increasing rotation speed would increase vortex depth and
the intensity of turbulent fluctuations on the free surface. He et al. [11] investigated the
coalescence and sedimentation behavior of falling droplets in a vessel of MSM. They found
that increasing rotation speed is conducive to the dispersion of droplets in the mechanical
stirring vessel. Later, they [12] also investigated the dispersed behavior of desulfurization
particles at different rotation speeds and found that rotation speed affects the liquid dis-
persion in the mechanical stirring vessel and has a noticeable effect on the distribution of
particles in the hot metal. Li et al. [13,14] used a Volume of Fluid with a dispersed particle
model (VOF-DPM) to analyze the drawdown mechanism and the dispersion characteristic
of light particles at different rotation speeds. They found that a higher impeller rotation
speed can provide more power consumption and increase particle velocity, which means
the mixing efficiency would be better. Nakai et al. [15,16] described the variation of the
gas–liquid interface at different stages of the mixing process in a 1/8 scale water model.
They found that the vortex height and operation parameters, such as impeller immersion
depths, rotation speeds, and the geometric size of KR, have a relationship. Ji et al. [17]
investigated the mixing behavior of KR vessels using computational fluid dynamics (CFD)
simulations. By comparing fixed- and variable-velocity stirring, they found the latter
stirring mode to be more favorable in the mixing of a bath. Sahu et al. [18] used CFD
simulation, which optimized the operating parameters in the synthesis of hybrid aluminum
matrix composites by stir casting. They found that blade angle, impeller size, and stirring
speed are the typical factors that affect the flow pattern. Liu et al. [19] investigated im-
peller immersion depths, rotation speeds, and eccentric mechanical stirring modes. They
found that increasing rotation speed can effectively strengthen the shear effect around
the impeller, which intensifies bubble disintegration and dispersion. With an increasing
impeller immersion depth, the distance between the vortex and impeller rises gradually.
All the related research above shows that the impeller immersion depth and rotation speed
significantly influence the fluid flow and mixing efficiency. However, to date, only a few
reports focus on this topic, and it is not fully understood or well documented due to the
limits of the measurement method and high costs of in-situ experimentation at a high
ambient temperature and within a hazardous environment.
This study aims to provide a detailed analysis of the effect of impeller operating
variables on fluid flow pattern, mixing characteristics, vortex core depth, and free surface
velocity of KRs based on the developed Volume of Fluid model with the sliding mesh
method (VOF-SMM). This study extends our previous work [20] in which the effect of the
Metals 2021, 11, 1596 3 of 18

impeller dimension was discussed. Here, our focus is the rotational speed and immersion
depth of impellers.

2. Mathematical Model
Our model has been reported in recent studies on the effect of impeller dimension for
the KR process [20] and on a new dephosphorization technology for the converter steelmak-
ing process [21]. In those studies, the grid sensitivity, calibration of the turbulence model,
and model validation have been carefully examined, but considering some improvements
and the completion of the present work, the outline is given as follows.

2.1. Governing Equations of VOF Model


To track the free surface of gas–liquid two-phase flow in a KR, some assumptions were
specified, and details were given in our previous work [20]. As a result, the hydrodynamics
of this multiphase system is governed by a set of single Navier–Stokes equations and a
continuity equation.
∂ρ ∂(ρui )
+ =0 (1)
∂t ∂xi

∂ρui ∂ ρui u j ∂p ∂τij
+ =− + + ρgi + f σ (2)
∂t ∂x j ∂xi ∂x j
Here, ui and uj are the velocity components in the direction of i and j, respectively
(m·s−1 );t is the time (s); p is the pressure (Pa); g is the gravity acceleration (m·s−2 ); and τ ij
is the Reynold stress tensor (N·m−2 ), which is calculated as follows:
!
∂u j ∂ui 2 ∂u
τij = µeff + − µeff k δij (3)
∂xi ∂x j 3 ∂xk

where δij is the Kronecker delta (δij = 1 if I = j and δij = 0 if i 6= j); µeff is the effective viscosity
(Pa·s) calculated by:
µeff = µ + µt (4)
where µ is the molecular viscosity (Pa·s) and the turbulent viscosity µt (Pa·s) is calculated
by the turbulence model.
In the present study, a modified approach was used based on the multi-fluid formula-
tion proposed by Ubbink and Isssa [22]. The evolution of the interface/free surface [23]
is described by an additional transport equation for the indicator function representing
the volume fraction of one phase that needs to be solved together with the continuity and
momentum equations:
 
∂αq ∂ ui α q ∂ ui,r αq 1 − αq
+ + =0 (5)
∂t ∂xi ∂xi

where, ui ,r = ui ,q − ui ,p is the relative velocity (m·s−1 ) between phase q and phase p.


Here, the third term on the left hand of Eqn. (5) is added to bring in a counter-gradient
artificial interface compression term in which the ui ,r ensures compression with
 a suitable
compression velocity, while the ∂/∂xi guarantees conservation and αq 1 − αq guarantees
boundedness with the compression only in the interface region.
In each cell, the volume fractions of all phases sum to unity, i.e.,
n
∑ αq = 1 (6)
q =1

The one-fluid formulation relies on the fact that multiple fluids (or phases) are not in-
terpenetrating. Because the same immiscible fluids are considered as an effective fluid
throughout the domain, the momentum equations shared by all phases are solved with
Metals 2021, 11, 1596 4 of 18

an effective density ρ (kg·m−3 ) and an effective viscosity µ (Pa·s). The fluid properties are
calculated by a weighted averaging method based on the volume fraction of each phase.
The volume-fraction-averaged density and viscosity are calculated, respectively, by:

ρ= ∑ αq ρq (7)
q

µ= ∑ αq µq (8)
q

In Equations (1)–(8), the subscript q denotes the liquid and gas phases.
To consider the effect of surface tension, an additional term is included in the momen-
tum equation as fσ on the interface S(t), calculated per unit volume using the continuum
surface force model with the following expression:
Z
fσ = σκ ∇αdS (9)
s(t)

where σ is surface tension (N·m−1 ) and κ is the mean curvature of the free surface deter-
mined by:
∇α
 
κ = −∇ · (10)
|∇α|

2.2. Turbulence Model


We have tested four kinds of classic turbulence models, i.e., the standard k-ε, standard
k-ε with swirl flow correction term, RNG k-ε, and SST k-ω, and found that the RNG k-ε
model can give an excellent solution compared with the experiments [20]. As a result, this
study employs the RNG k-ε model to close the turbulence. Thus, the turbulence kinetic
energy k equation is given by
" #
∂ ∂ ∂ ∂k
(ρk) + (ρui k) = (µ + µt ) + Gk + Gb − ρε (11)
∂t ∂xi ∂x j ∂x j

where Gk represents the generation of turbulent kinetic energy due to the mean velocity
gradients. Gb represents the generation of turbulent kinetic energy due to buoyancy.
Turbulent energy dissipation rate ε equation is given by
" #
∂ ∂ ∂ ∂ε ε ε2
(ρε) + (ρui ε) = (µ + µt ) + C1ε ( Gk + C3ε Gb ) − C2ε ρ (12)
∂t ∂xi ∂x j ∂x j k k
 
k
µt = µt0 f αs , Ω, (13)
ε
where µt is the turbulence viscosity, µt0 is the value of turbulent viscosity without swirl
modification, Ω is a characteristic swirl number, and αs is a swirl constant set as 0.07 in the
current study. The constants used here are C1ε = 1.44 and C2ε = 1.92.

2.3. Sliding Mesh Method for Impeller Motion


The simulation of KRs involves the impeller motion, which requires specified numeri-
cal tactics. In the current study, the sliding mesh method (SMM) [24–29] is adapted, and
full transient simulations are carried out using two grid zones (see Figure 1). Compared
with other technologies used to numerically deal with the region motion, such as the
multiple reference frame (MRF) method, there are no additional source terms inserted into
the momentum equations, e.g., centrifugal, Coriolis, and Euler forces, but the mesh will be
moving. In SMM, one grid zone (a rotational zone) is attached to the impeller to model the
rotating motion during the mechanically stirred process. In contrast, the other grid zone is
attached to the vessel wall (a fixed region), as shown in Figure 1. Since the nodes in the
full transient simulations are carried out using two grid zones (see Figure 1). Compared
with other technologies used to numerically deal with the region motion, such as the mul-
tiple reference frame (MRF) method, there are no additional source terms inserted into the
momentum equations, e.g., centrifugal, Coriolis, and Euler forces, but the mesh will be
Metals 2021, 11, 1596 moving. In SMM, one grid zone (a rotational zone) is attached to the impeller to model 5 of 18
the rotating motion during the mechanically stirred process. In contrast, the other grid
zone is attached to the vessel wall (a fixed region), as shown in Figure 1. Since the nodes
in rotating
the rotating
regionregion
movemove rigidly
rigidly in a given
in a given rotational
rotational zone
zone for for SMM,
SMM, it is imperative
it is imperative that the
that
interface between the zones consists of two overlapping faces, one of which of
the interface between the zones consists of two overlapping faces, one which is to
is attached
attached to theregion
the rotating rotating region
while thewhile
other the other is attached
is attached to the
to the fixed fixedSpecifically,
region. region. Specifically,
the sliding
theinterface
sliding consists
interfaceofconsists of a set of identical surface elements accessible
a set of identical surface elements accessible from both sides from ofboth
the in-
sides of the
terface. In interface. In a single
a single movement movement
step, the meshstep,
in thethe mesh region
moving in the moving
slides with region slides
a predefined
with a predefined
velocity velocity
across the mesh across the mesh
in the fixed in the
region. fixed
After region.
each step, After each step,
the interface the inter-
vertices in the
face vertices in the moving and static region will be reattached according
moving and static region will be reattached according to the initially computed vertex to the initially
map
computed
list. The vertex mapbetween
interaction list. The the
interaction between
two regions the twobyregions
is modeled is modeled
interpolating by inter-
the information
polating the information
across these across
interfaces. The SMMthese interfaces.
algorithm The SMM
is used algorithm
to allow is used
for velocity andto allow for
pressure data
velocity
transferand pressure
from data transfer
one domain from and
to the other one vice
domain to the other and vice versa.
versa.

Figure
Figure 1. The
1. The computational
computational domain
domain consisted
consisted of the
of the fixed
fixed andand rotational
rotational regions
regions andand separated
separated thethe
sliding mesh interface.
sliding mesh interface.

2.4. Tracer Transport Equation


2.4. Tracer Transport Equation
To measure the mixing performance of KRs, the mixing time as a significant indicator
To measure the mixing performance of KRs, the mixing time as a significant indicator
has been adapted. Consequently, a passive scalar transport equation of the tracer, which is
has been adapted. Consequently, a passive scalar transport equation of the tracer, which
essentially governed by a convective–diffusion equation, has to be solved at each time step.
is essentially governed by a convective–diffusion equation, has to be solved at each time
step.
" #
∂ ( αL C ) ∂ ( αL C ) ∂ ∂ ( αL C )
ρ + ρu j − Deff =0 (14)
𝜕(𝛼L 𝐶)∂t 𝜕(𝛼L 𝐶) ∂x j 𝜕 ∂x j 𝜕(𝛼 L 𝐶) j
∂x
𝜌 + 𝜌𝑢𝑗 − [𝐷 ]=0 (14)
𝜕𝑡 𝜕𝑥𝑗 𝜕𝑥𝑗 eff 𝜕𝑥𝑗
where C is the local tracer concentration, αL is the liquid phase volume fraction, and Deff is
the effective
where C is the diffusivity.
local tracerThe effective diffusion
concentration, αL is thecoefficient
liquid phasemayvolume
be calculated as the
fraction, andsum
Deff of
the molecular D and turbulent diffusivities D . Here, D is set as 10 −9 m2 /s, a typical
is the effective diffusivity.
lam The effective diffusion coefficient
t may be calculated as the sum
lam
of value for a solute
the molecular Dlaminand
liquids,
turbulent Dt is calculated
and diffusivities from D
Dt. Here, thelamturbulent
is set as 10kinetic
−9 m2/s,viscosity
a typical νt
as Dfor
value t = aνtsolute
/Sct , where Sct isand
in liquids, theDturbulent Schmidt
t is calculated fromnumber and itskinetic
the turbulent value 0.7 is used.
viscosity The
νt as
velocity field used in the above tracer equation was obtained from a steady solution of VOF
simulation and remained as a constant value during the calculation of this passive scalar.
The initial tracer concentrations are designated as zero everywhere in the vessel
except at the blue Region A, as shown in Figure 2. To monitor the variation of tracer with
time when measuring the mixing time, seven probe points were placed in several typical
locations. Their coordinates are also listed in Figure 2.
Dt = νt/Sct, where Sct is the turbulent Schimidt number and its value 0.7 is used. The veloc-
ity field used in the above tracer equation was obtained from a steady solution of VOF
ity field used in the above tracer equation was obtained from a steady solution of VOF
simulation and remained as a constant value during the calculation of this passive scalar.
simulation and remained as a constant value during the calculation of this passive scalar.
The initial tracer concentrations are designated as zero everywhere in the vessel ex-
The initial tracer concentrations are designated as zero everywhere in the vessel ex-
cept at the blue Region A, as shown in Figure 2. To monitor the variation of tracer with
cept at the blue Region A, as shown in Figure 2. To monitor the variation of tracer with
Metals 2021, 11, 1596 time when measuring the mixing time, seven probe points were placed in several typical 6 of 18
time when measuring the mixing time, seven probe points were placed in several typical
locations. Their coordinates are also listed in Figure 2.
locations. Their coordinates are also listed in Figure 2.

Figure 2. Schematic illustration of tracer addition region and the position of monitoring points with
Schematicillustration
Figure2.2.Schematic
Figure illustrationof
oftracer
traceraddition
additionregion
regionand
andthe
theposition
positionof
ofmonitoring
monitoring points
points with
with
detailed coordinates.
detailed
detailedcoordinates.
coordinates.

2.5.
2.5.Computational
ComputationalDetails
Detailsand
andBoundary
BoundaryConditions
Conditions
2.5. Computational Details and Boundary Conditions
AAmesh
meshwith
with450,000
450,000cells,
cells,which
whichis a reasonable
reasonable balance
balance between
between thethe computational
A mesh with 450,000 cells, which isisaareasonable balance between the computational
computational
cost
cost (630,000
(630,000 cells)
cells) and
and accuracy
accuracy (260,000
(260,000 cells),
cells), was
was used
used after
after the
the grid
grid independence
independence test
test
cost (630,000 cells) and accuracy (260,000 cells), was used after the grid independence test
in
in the previous
previous research,
research,which
whichshowed
showed that
that thethe change
change of of simulated
simulated results
results was waswhen
<4% <4%
in the previous research, which showed that the change of simulated results was <4%
when
furtherfurther refining
refining mesh.mesh. The typical
The typical meshmesh is shown
is shown in Figure
in Figure 3. Figures
3. Figure 3a,b show
3a,b show the
the mesh
when further refining mesh. The typical mesh is shown in Figure 3. Figures 3a,b show the
mesh
detailsdetails on the
on theoncentralcentral vertical
vertical sections
sections and the and the impeller.
impeller.
mesh details the central vertical sections and the impeller.

Numericalmesh
Figure3.3.Numerical
Figure meshof
of(a)
(a)central
centralvertical
verticalprofiles
profiles and
and (b)
(b) impeller.
impeller.
Figure 3. Numerical mesh of (a) central vertical profiles and (b) impeller.
As for boundary conditions in the current calculation, the top surface of the vessel is
treated as a pressure outlet boundary, and the ladle wall, bottom, shaft, and impeller are
treated as the no-slip wall boundary condition. The PISO algorithm is used to decouple
the pressure–velocity. The discretization of the convective terms in the momentum and
turbulence model equations employed the QUICK scheme, and the time step was set
to 0.001 s. The calculations were carried out using the commercial computational fluid
dynamic software FLUENT 14.5. The main convergence criterion is that the residuals for
torque, velocity, and pressure achieve a steady state, i.e., they do not change significantly
with further increasing time steps or iterations. The second criterion for checking the
Metals 2021, 11, 1596 7 of 18

residuals was committed for all governing equations to less than 1 × 10−5 . The detailed
geometric and physical parameters in this study are shown in Table 1.

Table 1. The geometrical, physical, and operating parameters used in this study.

Items Values
Ladle diameter, D (mm) 200
Ladle height, H (mm) 300
Blade length, d (mm) 80
Blade height, h (mm) 60
Blade thickness, w (mm) 20
Impeller immersion depth, I (mm) 100, 110, 120, 130
Impeller rotation speed, ω (rpm) 180, 200, 220, 240, 260
Height filled with liquid water, H2 (mm) 200
Water density, ρs (kg/m3 ) 998.2
Water dynamic viscosity, µs (Pa·s) 0.001
Top air density, ρo (kg/m3 ) 1.225
Top air dynamic viscosity, µo (Pa·s) 1.79 × 10−5
Total mesh number (-) ~450,000

2.6. Model Validation


Model validations have been reported [20]. Because the availability of the industrial
KR process data is limited by the measurement technology and the high costs of in-situ
experimentation due to a high temperature under hazardous conditions, the water model
experiments were resorted to in order to validate the model. Here, similar conditions of
the water model for KR have been adapted. In earlier studies, the effect of the impeller
dimension on the fluid flow and interface behavior was investigated based on numerical
simulation. It was found that the flow field and interface behavior of the process of KR hot
metal pretreatment can be represented by a 3D Volume of Fluid (VOF) model coupling the
sliding mesh method and RNG k-ε model. The water model experiment was designed to
validate the present mathematical model. The simulated results based on the current model
are in good agreement with the measured results in the water model, which corroborates
the applicability of the proposed model.
Furthermore, concerning the rotational speed and immersion depth, the comparison
of the vortex profile between a water model and numerical simulation at the rotation
velocity of 180 rpm and different impeller immersion depths is shown in Figure 4. Figure 5
shows the vortex depth of the water model and the numerical simulation at the impeller
immersion depth of 110 mm with different rotation speeds. It is observed that the shape
and vortex depth can be well captured, and specific agreement between the simulated8 and
Metals 2021, 11, x FOR PEER REVIEW of 19
experimental results can reasonably arrive.

Figure4.4.Comparison
Figure Comparisonof
ofvortex
vortexdepth
depthof
ofnumerical
numericalsimulation
simulationand
andwater
watermodel
modelat
atimpeller
impellerimmer-
immer-
sion depths of 100 mm and 120 mm.
sion depths of 100 mm and 120 mm.
Metals 2021, 11, 1596 8 of 18
Figure 4. Comparison of vortex depth of numerical simulation and water model at impeller immer-
sion depths of 100 mm and 120 mm.

Figure5.5.Comparison
Figure Comparisonof ofvortex
vortexdepth
depthof
ofnumerical
numericalsimulation
simulationand
andwater
watermodel
modelatatimpeller
impellerrotation
rotation
velocities of 220 rpm and 260 rpm.
velocities of 220 rpm and 260 rpm.

3.3.Results
Resultsand
andDiscussion
Discussion
3.1. Fluid Flow Pattern and Qualifying Inactive Zone
3.1. Fluid Flowflow
The bath Pattern and Qualifying
in KRs Inactive
is driven by Zone
the high-speed rotating impeller, which directly
Thekinetic
transfers bath flow
energyin KRs
to theisliquid
driven bymixes
and the high-speed rotating
it with reagent impeller,
powders which directly
for desulfurization.
transfers kinetic energy to the liquid and mixes it with reagent
Consequently, it is worth investigating the effect of impeller immersion depth powders for desulfuriza-
(IID) and
tion. Consequently,
impeller rotation speed it is(IRS)
worthoninvestigating
the stirring andthe effect
mixingof characteristic
impeller immersionof KRs.depth
Figure (IID)
6
and impeller
shows rotation
the central speed
vertical (IRS) on thestreamlines
cross-section stirring andandmixing characteristic
a close-view of theoflocal
KRs.region
Figure
6 shows
around the
the centralat
impeller vertical cross-section
an immersion depthstreamlines
of 100 mm, 110 andmm,
a close-view
120 mm, of the
and local
130 mm.region
The
around theconditions
simulation impeller at cananbeimmersion depth1,ofand
found in Table 100here
mm,the110 mm, 120
rotation speedmm,of and 130 mm.
the impeller
isThe simulation
fixed to 180 rpm. conditions
Similar to canourbe previous
found instudy,
Table two
1, and here
main the rotation
recirculation speed
flow of the
patterns,
where
impeller is fixed to 180 rpm. Similar to our previous study, two main recirculationflow
the flow field is separated into the upper and lower zones by the discharge flow
driven
patterns,by where
the impeller
the flow rotation, were also into
field is separated observed in Figure
the upper 6. With
and lower theby
zones increase of the
the discharge
impeller immersion
flow driven depth, the
by the impeller root location
rotation, were alsoof discharge
observedflow moves
in Figure 6. downward, but theof
With the increase
discharge
the impeller strength, i.e., the
immersion maximum
depth, the rootvelocity,
locationappears to remain
of discharge flow invariant for all cases;
moves downward, but
see
thethe corresponding
discharge strength, close-view of the localvelocity,
i.e., the maximum region around
appearsthe to impeller in Figure for
remain invariant 6a–d.all
Metals 2021, 11, x FOR PEER REVIEW
As a result, the upper recirculation zone was expanded. This flow pattern
cases; see the corresponding close-view of the local region around the impeller in Figure change 9
would of 19
influence the local and global mixing characteristics.
6a–d. As a result, the upper recirculation zone was expanded. This flow pattern change
would influence the local and global mixing characteristics.

Figure6.6.The
Figure Thestreamlines
streamlinesofofthe
thecentral
centralvertical
verticalcross-section
cross-sectionatatan
animmersion
immersiondepth
depthofof(a)
(a)100
100mm,
mm,
(b) 110 mm, (c) 120 mm, and (d) 130
(b) 110 mm, (c) 120 mm, and (d) 130 mm. mm.

In contrast, the flow pattern under different IRSs is also shown in Figure 7a–e, which
gives the vertical cross-section flow field and streamlines at IRSs of 180 rpm, 200 rpm, 220
rpm, 240 rpm, and 260 rpm. Flow velocity in the bath, both local and global, has been
increased significantly with the increase of IRS. However, the root location of discharge
Figure 6. The streamlines of the central vertical cross-section at an immersion depth of (a) 100
Metals 2021, 11, 1596 9 of 18
(b) 110 mm, (c) 120 mm, and (d) 130 mm.

In contrast, the flow pattern under different IRSs is also shown in Figure 7a–e, w
gives
In the vertical
contrast, cross-section
the flow flow
pattern under field and
different IRSsstreamlines
is also shownatinIRSs
Figureof7a–e,
180 rpm,
which200 rpm
gives the vertical cross-section flow field and streamlines at IRSs
rpm, 240 rpm, and 260 rpm. Flow velocity in the bath, both local and global, of 180 rpm, 200 rpm, has
220 rpm, 240 rpm, and 260 rpm. Flow velocity in the bath, both local and global, has been
increased significantly with the increase of IRS. However, the root location of disch
increased significantly with the increase of IRS. However, the root location of discharge flow
flowincreasing
with with increasing
IRS showsIRS shows
almost almost
no change. no change.
These These
results reveal thatresults reveal
the effects that
of IRS andthe effec
IRSonand
IID the IID
flowon the flow
pattern pattern These
are different. are different. These
results also makeresults alsotomake
it possible it possible to
operationally
ationally
control control
the flow the in
pattern flow pattern
a KR via the in a KR via the complementarity
complementarity of the two parameters. of the two param

Figure 7. The flow field and streamlines of the central vertical cross-section at impeller rotation
Figure 7. The flow field and streamlines of the central vertical cross-section at impeller ro
speeds (IRSs) of (a) 180 rpm, (b) 200 rpm, (c) 220 rpm, (d) 240 rpm, and (e) 260 rpm.
speeds (IRSs) of (a) 180 rpm, (b) 200 rpm, (c) 220 rpm, (d) 240 rpm, and (e) 260 rpm.
Further, some quantities comparisons were also conducted. Figure 8a,b show the axial
Further,
and radial some
velocity quantities
variations alongcomparisons
Lines A and B.were
Line Aalso
is a conducted.
vertical line ofFigures
0.075 m8(a),
from (b) show
axial and radial velocity variations along Lines A and B. Line A is a vertical line of
the axial centerline of the vessel, and Line B is a horizontal line of 0.85 m from the bottom
surface,
m fromwhich is illustrated
the axial in the
centerline of subgraph
the vessel,of and
Figure 8a(1).
Line B isThrough comparison,
a horizontal line ofit 0.85
can m from
be seen that the changing trend of velocities on Line A is the same. However, the level of
bottom surface, which is illustrated in the subgraph of Figure 8 (a1). Through compar
maximum velocity decreases with the increase of IID, which can also be observed from
the velocity contour of the central vertical section at different IIDs. Figure 8a(2) shows that
the radial velocity distribution is almost axis-symmetrical. The velocity near the outside
edge of the impeller is greater than at the center and near the wall. It can also be seen
in Figure 8a(2) that the velocity along Line B below the impeller slightly increases when
increasing IID, and the closer it is to the edge of the impeller, the more pronounced this
increase is. Besides, the velocity below the impellers is always minimal, reaching almost
zero. Finally, it should be stressed that the velocity change caused by the IID change is
finite compared with the influence of IRS.
Figure 8b(1,2) show the axial and radial velocity variations at different IRSs. It can be
observed that the axial and radial velocity increases significantly, and the diameter of the
lower-velocity region decreases as the impeller rotation speed increases. However, for all
the impeller rotation velocities, the velocity below the impeller is still miniscule, making it
particularly difficult to achieve global mixing of the species in the reactor. Furthermore,
Figure 9 shows the velocity distribution contour of horizontal sections with different heights
under the different IRSs. It can be seen that the liquid velocity of different horizontal
sections increases noticeably with increasing IRS. However, the velocity below the impeller
is still very small no matter how fast the IRS is. These phenomena were also observed by T.
Yamamoto et al. [7–10] in the flow characteristic of aluminum melt by mechanically stirring.
miniscule, making it particularly difficult to achieve global mixing of the species in the
reactor. Furthermore, Figure 9 shows the velocity distribution contour of horizontal sec-
tions with different heights under the different IRSs. It can be seen that the liquid velocity
of different horizontal sections increases noticeably with increasing IRS. However, the ve-
locity below the impeller is still very small no matter how fast the IRS is. These phenomena
Metals 2021, 11, 1596 10 of 18
were also observed by T. Yamamoto et al. [7–10] in the flow characteristic of aluminum
melt by mechanically stirring.

Metals 2021, 11, x FOR PEER REVIEW 11 of 19


Figure8.8.Comparison
Figure Comparison ofof
velocity distributions
velocity along
distributions alongwith Line
with A for
Line thethe
A for axial direction
axial in sub-graph
direction in sub-
(1)graph
and Line B forLine
(1) and the B
radial direction
for the in sub-graph
radial direction (2) at different
in sub-graph (a) IIDs and
(2) at different (a) (b)
IIDsIRSs.
and (b) IRSs.

The velocity
Figure 9. The velocity distribution
distribution of
of different
different levels
levels at
at IRS
IRS of
of (a)
(a) 180
180 rpm,
rpm, (b)
(b) 200
200 rpm,
rpm, (c)
(c) 220
220 rpm,
rpm,
(d) 240 rpm and (e) 260 rpm.

To further study the effect of operation parameters on the bath flow of KRs, the mean
velocity of the fluid in the bath under different IIDs was also investigated by indirectly
measuring the dynamic energy of the bath obtained from the mechanical stirring. Figure
10a shows the comparison of the mean velocities for different IIDs. As expected, when the
immersion depth increases, the mean velocity is basically unchanged. The comparison of
the mean velocity of the bath under different IRSs is shown in Figure 10b. The mean ve-
Metals 2021, 11, 1596 11 of 18

Figure 9. The velocity distribution of different levels at IRS of (a) 180 rpm, (b) 200 rpm, (c) 220 rpm,
(d) 240 rpm and (e) 260 rpm.

To further study the effect of operation parameters on the bath flow of KRs, the mean
To further study the effect of operation parameters on the bath flow of KRs, the mean
velocity
velocity ofof the
the fluid
fluidininthe
thebath
bathunder
underdifferent
differentIIDs
IIDs was
was also
also investigated
investigated by by indirectly mea-
indirectly
suring the dynamic energy of the bath obtained from the mechanical
measuring the dynamic energy of the bath obtained from the mechanical stirring. Figure stirring. Figure 10a
shows
10a shows the comparison of the mean velocities for different IIDs. As expected, when thewhen the
the comparison of the mean velocities for different IIDs. As expected,
immersion
immersion depth depth increases,
increases, the the
mean mean velocity
velocity is basically
is basically unchanged.
unchanged. The comparison
The comparison of
of the mean velocity of the bath under different IRSs is shown in Figure 10b.ve-
the mean velocity of the bath under different IRSs is shown in Figure 10b. The mean The mean
locity increases
velocity increaseslinearly as the
linearly asrotation speedspeed
the rotation increases but remains
increases invariant
but remains when IIDwhen IID
invariant
changes. The
changes. Themeanmean velocity
velocityof the liquid
of the risesrises
liquid 0.0220.022
m/s as the as
m/s impeller rotationrotation
the impeller velocity velocity
increases by 10 rpm. Unlike related research, the relationship between
increases by 10 rpm. Unlike related research, the relationship between the mean the mean velocity
velocity of
of the bath and IRS has been further obtained. A fitting regression, where the mean veloc-
the bath and IRS has been further obtained. A fitting regression, where the mean velocity v
ity 𝑣̅ is correlated with the impeller rotation speed ω, is calculated as follows:
is correlated with the impeller rotation speed ω, is calculated as follows:
𝑣 = −0.0009 + 0.0022𝜔 (180 ≤ 𝜔 ≤ 260) 𝑅2 = 0.9976 (15)
v = −0.0009 + 0.0022ω (180 ≤ ω ≤ 260) R2 = 0.9976 (15)

Metals 2021, 11, x FOR PEER REVIEW 12 of 19


Figure Comparison
10.Comparison
Figure 10. of theofmean
the velocities
mean velocities of inthe
of the fluid thefluid indifferent
bath at the bath (a) at different
impeller (a) impeller
immer-
sion depths (IIDs)
immersion depths and (b) IRSs.
(IIDs) and (b) IRSs.
Additionally,totofurther
Additionally, furtherexamine
examinethe the effect
effect of of
IIDIIDandand
IRSIRS
on on
thethe volume
volume fraction
fraction of
of the
the inactive
inactive zone,zone, an indicator
an indicator γ is defined,
γ is defined, whichwhich is theofratio
is the ratio of thesize
the region region sizevelocity
of flow of flow
velocity when it is less than 0.1 m/s to the fluid volume in the bath. The γ at different
when it is less than 0.1 m/s to the fluid volume in the bath. The γ at different IIDs is shown IIDs
is shown in Figure 11. With the increase of IID, the γ decreased from
in Figure 11. With the increase of IID, the γ decreased from 7.63% to 4.92%. Therefore, it7.63% to 4.92%.
Therefore,
can it canthat
be inferred be the
inferred thatimmersion
impeller the impeller immersion
depth depth does affect
does significantly significantly affect
local mixing.
local mixing. A fitting regression of the volume fraction of the inactive zone
A fitting regression of the volume fraction of the inactive zone γ as a function of impellerγ as a function
of impeller depth
immersion immersion depth I as
I is obtained is obtained
follows. as follows.
𝛾 = 36.646 − 0.4425𝐼 + 0.0015𝐼 2 2 (100 ≤ 𝐼 ≤ 130) 𝑅22 = 0.9980 (16)
γ = 36.646 − 0.4425I + 0.0015I (100 ≤ I ≤ 130) R = 0.9980 (16)
Volume fraction of inactive zone (%)

8 7.63

6.49
6
5.45
4.92

0
100 110 120 130
Immersion depth (mm)

Figure 11.
Figure 11. Comparison
Comparison of
of the
the volume
volume fraction
fraction of
of inactive
inactive zones
zones at
at different
differentIIDs.
IIDs.

The volume fraction of the inactive zone at different rotation velocities is shown in
Figure 12. The γ of the stirred molten pool decreases from 6.49% to 4.32% as the impeller
rotation velocity increases from 180 rpm to 260 rpm. The γ decreases by 0.92% as the im-
peller rotation velocity increases from 180 rpm to 200 rpm, which is the maximum decline
Volume
2

0
100 110 120 130
Metals 2021, 11, 1596 Immersion depth (mm) 12 of 18

Figure 11. Comparison of the volume fraction of inactive zones at different IIDs.

Thevolume
The volumefraction
fractionof ofthe
the inactive
inactive zone
zone at
at different
different rotation
rotation velocities
velocities is is shown
shown in in
Figure12.
Figure 12.TheTheγγof
ofthe
thestirred
stirredmolten
moltenpool
pooldecreases
decreasesfromfrom6.49%
6.49%to to 4.32%
4.32% asas the
the impeller
impeller
rotationvelocity
rotation velocityincreases
increasesfromfrom180180rpm
rpm toto 260
260 rpm.rpm.TheThe γ decreases
γ decreases by 0.92%
by 0.92% as theas im-
the
impeller rotation velocity increases from 180 rpm to 200 rpm, which
peller rotation velocity increases from 180 rpm to 200 rpm, which is the maximum decline is the maximum
decline
in in the
the range ofrange of 180
180 rpm rpmrpm.
to 260 to 260
Therpm. The γlinearly
γ drops drops linearly and monotonically
and monotonically in this
in this range,
range,
i.e., the i.e., the γ decreases
γ decreases by 0.2% by 0.2%
as the as the impeller
impeller rotation increases
rotation velocity velocity increases
by 10 rpm. byFurther,
10 rpm.
aFurther, a fitting regression
fitting regression of γ as aoffunction
of γ as a function impeller of impeller
rotation speedrotation speed ωas
ω is obtained is follows:
obtained
as follows:
2
𝛾 = 9.63
γ =−9.63
0.02𝜔 − 0.02ω (200(200 ≤ ω 𝜔≤ ≤ 260)
260) 𝑅 R2 = = 0.9889
0.9889 (17)
(17)

Figure 12. Variation of the volume fraction of the inactive zones at different IRSs.
Figure 12. Variation of the volume fraction of the inactive zones at different IRSs.
3.2. Mixing Time
Mixing time is an essential index to measure mixing characteristics. In order to
obtain the mixing time of the reactor, the stimulus–response method was used here. When
simulations/experiments are fully developed, a certain amount of tracer, specifically, 50 mL
of saturated KCl solution, is added to the reactor. The change of tracer concentration is
recorded, and the mixing time defined as the time span from the moment the tracer is
supplied to that when the concentration curves of all monitoring points approximate a
constant value with a fluctuation range restricted to ±5 pct. Some related computational
details are given in Section 2.4. Figure 13 plots the specific tracer concentration change at
those monitoring points defined in Figure 2. Here, the response curves of tracer with time
under different IRSs are shown in Figure 13a–e, which correspond to the IRSs of 180 rpm,
200 rpm, 220 rpm, 240 rpm, and 260 rpm, respectively. In Figure 13, the red vertical line
indicates the mixing time based on the above definition. The curve of each point shows
how to respond to the change of tracer concentrations and reveals the flow pattern to a
certain extent. For example, monitoring point 7, the curves converge gradually on the
equilibrium concentration for IRSs of 180, 200 and 220 rpm. Next, the curve has one peak
for the IRS at 240 rpm. As the IRS increases to 260 rpm, the curves quickly converge on the
equilibrium concentration; moreover, the period shortens with the rise of the IRS. Since the
rotating impeller drives the fluid flows, the rotating velocity directly impacts the kinetic
energy transferred to the fluid. Thus, with the increase of IRS, the response of the tracer to
the fluid hydrodynamic becomes more and more intense, and the tracer concentration at
each monitoring point arrives more quickly at the theoretical equilibrium concentration,
which suggests the mixing should be reinforced significantly with increasing IRS.
Metals Metals 2021,
2021, 11, 11, 1596
x FOR PEER REVIEW 13 of 18
14 of 19

Figure 13. The


Figure change
13. The changeof of
tracer
tracerconcentration ofmonitoring
concentration of monitoring points
points with
with timetime
at IRSatofIRS of (a)
(a) 180 180
rpm, (b)rpm, (b) 200
200 rpm, rpm,
(c) 220 (c) 220
rpm,
rpm, (d)
(d) 240
240rpm,
rpm,and
and(e)(e)
260260 rpm.
rpm.

Furthermore,the
Furthermore, themixing
mixing time
time at
atdifferent
differentIRSs is shown
IRSs in Figure
is shown 14. It14.
in Figure decreases
It decreases
significantly with the increase of IRS. As can be seen from the above, with the increase of
significantly with the increase of IRS. As can be seen from the above, with the increase of
IRS, the mean velocity of the bath increases significantly. The minimum mixing time is 55 s
IRS, the mean velocity of the bath increases significantly. The minimum mixing time is 55
for the IRS at 260 rpm—a maximum value in this study’s range. Meanwhile, the volume
s for the IRS
fraction of at
the260 rpm—a
inactive zonemaximum value inwhich
decreases visibly, this study’s
producesrange. Meanwhile,
a decrease in mixing the volume
time.
fraction of the inactive
Consequently, the rise zone
of thedecreases visibly,means
IRS is an effective whichtoproduces a decrease
shorten the in mixing time.
mixing time.
Consequently, the rise of the IRS is an effective means to shorten the mixing time.
Metals 2021, 11, x FOR PEER REVIEW 15 of 19
Metals
Metals 2021,
2021, 11,
11, x1596
FOR PEER REVIEW 1514of
of 19
18

Figure 14.Comparison
Figure Comparison ofmixing
mixing timesatatdifferent
different IRSs.
Figure 14.
14. Comparison of
of mixing times
times at different IRSs.
IRSs.

Themixing
The
The mixingtime
mixing timeunder
time underdifferent
under differentIIDs
different IIDsis
IIDs isshown
is shownin
shown inFigure
in Figure15.
Figure 15.ItIt
15. Itcan
canbe
can beseen
be seenthat
seen thatthe
that the
the
variation
variation trend
variation trend of
trendofofmixing mixing
mixingtime time
time with with
withIIDIIDIID is
is first first
reduced
is first reduced
and then
reduced and then
and remains remains almost
almost almost
then remains unchanged. un-
un-
changed.
The minimum
changed. Theminimum
The minimum
mixing time mixing
mixingis 71.5time
times atisisan71.5
IIDssof
71.5 at130
at anIID
an IIDof
mm. ofPractically,
130mm.
130 mm.Practically,
Practically,
when the IID when
when the
range
the
IID range
is between
IID is
range is 110 between
mm and
between 110 mm
110 130
mm mm, and
and 130130
it alwaysmm, it
mm, ittakes always
alwaysabouttakes
takes about
72about 72
s to achieve s to achieve global
global mixing.
72 s to achieve global
mixing. Obviously, IID has a limited impact on mixing time
mixing. Obviously, IID has a limited impact on mixing time in this IID range. As isabove,
Obviously, IID has a limited impact on mixing time in this IID in this
range. IID
As isrange. As
indicated is indi-
indi-
catedthe
with
cated above,
increase
above, withofthe
with the
IID,increase
the mean
increase ofIID,
of IID,the
velocitythemeanmean
of velocity
thevelocity
bath hasof ofthe
thebath
almost bath hasalmost
no change,
has almost nochange,
but the
no change,
volume
but the
fraction volume
of the fraction
inactive of
zone the inactive
decreases zone
visibly. decreases
The volumevisibly.
but the volume fraction of the inactive zone decreases visibly. The volume fraction of The
fraction volume
of the fraction
inactive ofthe
zone the
at
inactive
an IID of zone
100 mmat an
is IID of 100
significantly mm is
greater significantly
than the othergreater
IIDs,
inactive zone at an IID of 100 mm is significantly greater than the other IIDs, making than
makingthe other
global IIDs,
mixing making
more
global mixing
global mixing
difficult more difficult
and increasing
more difficult
the mixing and increasing
and increasing
time. the mixing
Consequently,
the mixing time. Consequently,
compared
time. Consequently,
with IID, IRS has compared
a more
compared
with IID,
significant IRS has
effect ona more
the significant
mixing effect
performance on ofthe
themixing
bath inperformance
with IID, IRS has a more significant effect on the mixing performance of the bath in KRs.KRs. of the bath in KRs.

Figure15.
Figure
Figure 15.Comparison
15. Comparisonof
Comparison ofmixing
of mixingtimes
mixing timesat
times atdifferent
at differentIIDs.
different IIDs.
IIDs.

3.3. Vortex
3.3.Vortex Core
VortexCore Depth
CoreDepth
Depth
3.3.
Vortex characteristics
Vortexcharacteristics
characteristicshavehavean
have animportant
importantinfluence
influence ononthethe dispersion
dispersion behavior
behavior ofpar-of
par-
Vortex
particles and the refining an important
efficiency, as reportedinfluence
in some on the dispersion
studies [20]. It behavior
can be seenof that
ticlesand
ticles andthetherefining
refiningefficiency,
efficiency,asasreported
reportedin insome
somestudies
studies[20].
[20].ItItcan
canbebeseen
seenthat
thatthere
there
there
is is almost
almost no no change
change to theto the profile
profile of the of the gas–liquid
gas–liquid interface
interface and and vortex
vortex core core depth
depth at dif-
is
atalmost
differentnoimpeller
change to the profiledepths,
immersion of the gas–liquid
as shown interface
in Figure andThe
16. vortex core depth
distance between at dif-
the
ferent
ferent impeller
impeller immersion
immersion depths,
depths, as shown
as shown in Figure 16. The distance between the bot-
bottom
tomof of the
ofthe
the vortex
vortex core
core andand
thethe
toptop edge
edge ofinthe
ofthe
the
Figure
impeller
16. The
impeller increases
distance
increases between the
continuously
continuously asbot-
asIID
IID IID
in-
tom
increases. vortex
However, core and
this the
doesn’ttop edge
mean of
that the impeller
immersion increases
depth continuously
of the impeller asdoes in-
not
creases. However,this
creases. thisdoesn’t
doesn’tmean
meanthat
thatthetheimmersion
immersiondepth depthof ofthe
theimpeller
impeller doesnot not
affect theHowever,
behavior of the gas–liquid interface; this is discussed further in Sectiondoes
3.4.
affect the behavior of the gas–liquid interface; this is discussed further
affect the behavior of the gas–liquid interface; this is discussed further in Section 3.4. in Section 3.4.
Metals 2021, 11, x FOR PEER REVIEW 16 of 19
Metals 2021, 11,
Metals 2021, 11, 1596
x FOR PEER REVIEW 15of
16 of 18
19

Figure 16. Vortex depth and gas–liquid interface profile at different IIDs with IRSs of 180 rpm.
Figure 16.
Figure Vortex depth
16. Vortex depth and
and gas–liquid
gas–liquid interface
interface profile
profile at
at different
different IIDs
IIDs with
with IRSs
IRSs of
of 180
180 rpm.
rpm.
Increasing rotation speed is an effective way to increase the vertex core depth (VCH),
Increasing rotation speed is an effective way to increase the vertex core depth (VCH),
Increasing
favoring rotation speed
the entrainment and is an effective
dispersion wayreagent
of the to increase
by an theimpeller.
vertex coreThedepthvortex(VCH),
depth
favoring the entrainment and dispersion of the reagent by an impeller. The vortex depth
favoring the entrainment and dispersion of the reagent by an impeller.
and gas–liquid interface profile under different rotation speeds are shown in Figure 17. The vortex depth
and gas–liquid interface profile under different rotation speeds are shown in Figure 17.
and
Thisgas–liquid
figure vividlyinterface
depictsprofile
how under different
impeller rotationrotation speeds are
speed strongly shown
impacts the invortex
Figurecore 17.
This figure vividly depicts how impeller rotation speed strongly impacts the vortex core
This
depth figure vividly depicts
and gas–liquid how profile.
interface impellerTherotation
vortexspeed
core strongly impacts increases
depth gradually the vortexascore the
depth and gas–liquid interface profile. The vortex core depth gradually increases as the
depth and gas–liquid
IRS increases. interface
Until reaching profile. The
a rotation speed vortex
of 260corerpm,depth gradually
the bottom increases
of the as the
vortex passes
IRS increases. Until reaching a rotation speed of 260 rpm, the bottom of the vortex passes
IRS
the increases. Until reaching
upper interface a rotation
of the impeller, speed
which of 260 rpm, thetobottom
is advantageous of the vortex
the dispersion behaviorpasses of
the upper interface of the impeller, which is advantageous to the dispersion behavior of
the upper
particles and interface
andthe of
therefiningthe impeller,
refiningefficiency.
efficiency. which
The is advantageous
results to the dispersion behavior of
particles The results of of
thethe vortex
vortex core core
depthdepth
withwith respect
respect to theto
particles
the and the
numerical refining efficiency.
simulation and water The results
model of the vortex
experiment under core depth rotation
different with respect speeds to
numerical simulation and water model experiment under different rotation speeds were
the
were numerical
also plottedsimulation
in Figure and17. water model experiment
T. Yamamoto et al. [7,8] under
have different
reported therotation speeds
morphological
also plotted in Figure 17. T. Yamamoto et al. [7,8] have reported the morphological changes
were
changesalso toplotted
a free in Figure
surface 17.
with T. Yamamoto
different IRSs. et al.
The [7,8]
present have reported
work
to a free surface with different IRSs. The present work further analyses the relationshipfurther the morphological
analyses the rela-
changes
tionship to a free
between surface
the with
vortex different
depth and IRSs.
impellerThe present
rotation work
speed.
between the vortex depth and impeller rotation speed. To further investigate the effect further
To analyses
further the
investigate rela-
the
tionship
effect
of IID ofandbetween
IIDIRSandon the
IRS vortex depth
theoninterfacial
the interfacialand impeller
behavior
behavior rotation
of KRs,
of KRs, speed.
the free
the free To further
surface
surface investigate
velocity
velocity the
is reported
is reported in
effect
Section 3.4. A fitting regression of the vortex core depth HVCD as a function of reported
in of
Section IID and
3.4. A IRS on
fitting the interfacial
regression of behavior
the vortex of KRs,
core the
depth free
H surface
VCD as a velocity
function is
of impeller
impeller
in Section
rotation
rotation 3.4. A
speed
speed ωωfitting regression
is obtained
is obtained of the vortex core depth HVCD as a function of impeller
as follows.
as follows.
rotation speed ω is obtained as follows.
𝐻VCD = −33.25 + 0.36𝜔 (180 ≤ 𝜔 ≤ 260) 𝑅2 = 0.9965 (18)
HVCD = −33.25 + 0.36ω (180 ≤ ω ≤ 260) R2 2= 0.9965 (18)
𝐻VCD = −33.25 + 0.36𝜔 (180 ≤ 𝜔 ≤ 260) 𝑅 = 0.9965 (18)

70 Simulation value
70 Fitted linear
Simulation value
60 Experimental
Fitted linear value
60 Experimental value
(mm)

50
(mm)

50
depth

40
depth

40
30
Vortex

30
Vortex

20
20
10
10
0
0
-10
180 200 220 240 260
-10
180 Impeller
200 rotation
220 velocity240
(rpm) 260
Impeller rotation velocity (rpm)
Figure 17.
Figure 17. Vortex
Vortex depth
depth and
and gas–liquid
gas–liquidinterface
interfaceprofile
profileat
atdifferent
differentIRSs
IRSswith
withIIDs
IIDsofof110
110mm.
mm.
Figure 17. Vortex depth and gas–liquid interface profile at different IRSs with IIDs of 110 mm.
Metals 2021, 11, 1596 16 of 18

Metals 2021, 11, x FOR PEER REVIEW 17 of 19

Metals 2021, 11, x FOR PEER REVIEW 3.4. Free Surface Velocity 17 of 19

The velocity on the gas–liquid interface strongly influences the entrainment and
3.4. Free Surface Velocity
entrapment of the desulfurization particles. The interface velocity at different IIDs is shown
3.4. The
Free velocity
Surface on the gas–liquid interface strongly influences the entrainment and en-
in Figure 18. Velocity
It clearly indicates that the velocity around the vortex core is much bigger
trapment
The of the
velocity desulfurization
on walls,
the gas–liquid particles.
interface The interface
strongly velocity theat different IIDs andis shown
than that
in Figureof18.
near the
Itdesulfurization
clearly indicates
i.e., there is a gradient
that theThe velocity
ofinfluences
around
velocity along
the vortex
entrainment
the radial
core is much
en-
direction.
bigger
It also
trapment
suggests the
that the regents particles.
around the vortexinterface
core velocity
are much at different
more activeIIDs is
than shown
those near the
inthan that18.
Figure near the walls,
It clearly i.e., there
indicates that is a gradient
the of velocity
velocity around the along
vortexthecore radial
is muchdirection.
biggerIt
walls.
alsothat Furthermore,
suggests thatwalls, with
the regents the increase
around of IID,
the vortex the distance between the bottom of the vortex
than near the i.e., there is a gradient of core are along
velocity much the more active
radial than those
direction. It
core
near and
the impeller
walls. increases,
Furthermore, withresulting
the increasein theof rotational
IID, the impeller’s
distance between effect
the on theofinterface
bottom
also suggests that the regents around the vortex core are much more active than those
velocity
the the
near vortex around
walls. andthe
coreFurthermore, bottom
impeller ofthe
theincrease
increases,
with
vortex
resultingcore inbeing
of IID, the less pronounced.
therotational impeller’s
distance between
Consequently,
theeffect
bottom on of
the the
velocity
interface gradient
velocity between
around the the axis–centerline
bottom of the vortex and
corewalls
being
the vortex core and impeller increases, resulting in the rotational impeller’s effect on the becomes
less insignificant
pronounced. Conse- when the
quently,
impeller the
is velocity
taken gradient
down. For between
example, the axis–centerline
the velocity
interface velocity around the bottom of the vortex core being less pronounced. Conse-on and
the walls becomes
gas–liquid insignificant
interface around the
when
vortex the
coreimpeller
is is
obviously taken down.
greater For
than example,
those nearthe velocity
the walls
quently, the velocity gradient between the axis–centerline and walls becomes insignificant on
for the gas–liquid
immersion interface
depths of 100 mm,
around
i.e.,
when thethe
the vortexofis
change
impeller core is obviously
interface
taken down. greater
velocity
For along
example,thanwith
those near the
radial
the velocity walls
the for
direction
on immersion
is relatively
gas–liquid depths
steep for the
interface
of
around100
small IID mm, i.e.,
the vortex the
(see Figure change
core is 18a). of
obviously interface
However,
greaterforvelocity
thanthe along
case
those with
of IID
near radial
at 130
the walls direction
formm, is relatively
the change
immersion depthsof velocity,
steep
ofalong for the
100 mm, small
the IID (see Figure 18a). However, for the caseradial
of IIDdirection
at 130 mm, the change
withi.e.,radial change
direction, of interface velocity
is relatively evenalong(seewith
Figure 18d). As aisresult,
relatively
the interface
of
steep velocity,
for the along
small with radial direction, is relatively even (see Figure 18d). As a result, the
velocity near theIID (see Figure
vortex 18a). However,
core decreases for the case
significantly with of IID
theat 130 mm, the
increasing IID, change
which reveals
interface
ofthat
velocity, velocity near
along with radial the vortex core decreases significantly with the increasing IID,
which the entrainment
reveals that the the anddirection,
dispersion
entrainment
is relatively
effect ofeven
anddecreases
dispersion the (see Figure
vortex
effect of thewith
18d). As by
isvortex
affected a result,
isincreasing
affected theby the
arrangement
the
interface
of IID. velocity near vortex core significantly the IID,
arrangement of IID.
which reveals that the entrainment and dispersion effect of the vortex is affected by the
arrangement of IID.

Figure 18.
Figure 18. Comparison
Comparison of of
gas–liquid interface
gas–liquid velocity
interface at different
velocity IIDs of IIDs
at different (a) 100
ofmm, (b) mm,
(a) 100 110 mm,
(b) 110 mm,
(c) 120 mm
(c) 12018.
mm and (d) 130
and (d) 130mm.
mm.
Figure Comparison of gas–liquid interface velocity at different IIDs of (a) 100 mm, (b) 110 mm,
(c) 120 mm and (d) 130 mm.
In
In addition,
addition, thethe
velocity distribution
velocity on the on
distribution gas–liquid interface at
the gas–liquid different at
interface IRSs was
different IRSs
investigated, and the simulation results are shown in Figure 19. The velocity on the gas–
wasIninvestigated,
addition, the velocity
and thedistribution
simulationonresults
the gas–liquid
are shown interface at different
in Figure 19. TheIRSs was
velocity on the
liquid interface
investigated, and increases
the simulationsignificantly are
withshown
the rise of IRS; 19.
thisThe
phenomenon oncan
the be ex-
gas–liquid interface increasesresults
significantly in Figure
with the rise of velocity
IRS; this phenomenon gas– can be
plained
liquid in terms
interface of a
increases high-speed rotating
significantly with impeller
the rise that intensely
of IRS;that powers
this intensely
phenomenon the velocity
can be of
ex-velocity
explained in
the gas–liquid
terms of a high-speed rotating impeller powers the of
plained
the in termsinterface.
gas–liquid interface.
Notably,rotating
of a high-speed Notably,
for the case
for the
of IRS
impeller
case of
at intensely
that 260 rpm, the
IRS at 260
bottom
powers
rpm, the
of the vortex
thebottom
velocity ofofthe vortex
core
the sunk below
gas–liquid the top
interface. edge offor
Notably, thethe
impeller.of Thus,
IRS at it260
makes the theinterface
bottom ofvelocity near
core
the sunk below
impeller sharply theincrease
top edge
and thecase
ofdramatically
impeller. Thus,rpm,
promotes it makes
the dispersion
the vortex
the interface
behavior velocity
of the near
core sunk below the top edge of the impeller. Thus, it makes the interface velocity near
the impeller
reagents and sharply increase
desulfurization and
efficiency. dramatically promotes the dispersion behavior of the
the impeller sharply increase and dramatically promotes the dispersion behavior of the
reagents and desulfurization efficiency.
reagents and desulfurization efficiency.

Figure 19. Comparison of gas–liquid interface velocity at different IRSs at (a) 180 rpm, (b) 200 rpm,
(c) 220 rpm, (d) 240 rpm, and (e) 260 rpm.
Metals 2021, 11, 1596 17 of 18

4. Conclusions
Based on the developed 3D VOF-SMM model, the influences of operating parameters
on bath flow, mixing characteristics, and interface behaviors for KR process are numerically
investigated and arrive at the following results:
• Impeller immersion depth and rotation speed have different effects on the fluid
flow pattern of the bath under the present study range. The root location of the
discharge flow moves downward with the impeller immersion depth increasing, but
the discharge strength and the mean velocity of the bath show hardly any change.
Comparatively, the increase in the impeller rotation speed significantly improves the
mean velocity, but there is little change in the position of discharge flow. Furthermore,
increasing impeller immersion depth or rotation speed can effectively reduce the
volume fraction of the inactive zone, but it cannot eliminate it. As a result, the
correlation equations of γ as a function of ω or I are formulated under the range of
180 ≤ ω ≤ 260 and 100 ≤ I ≤ 130, respectively.
• Increasing impeller rotation speed is the most direct and effective way to shorten the
mixing time. However, the impeller immersion depth has a limited impact on the
mixing time by comparison. In the present study, a minimum mixing time of 55 s is
achieved at the maximum impeller rotation speed of 260 rpm.
• The vortex core depth and the velocity at the gas–liquid interface increase significantly
with the increasing impeller rotation speed, and a linear fitting regression has been
proposed. However, while the impeller immersion depth has little effect on the vortex
core depth, it has a visible influence on the velocity distribution of the free surface.
The velocity gradient on the gas–liquid interface between the axis to walls becomes
steep with the decreasing impeller immersion depth.

Author Contributions: Q.L., Z.Z. and M.L. conceived and designed the study. X.S. and S.M. accom-
plished the numerical simulation and data arrangement. The edition work was organized by Q.L.
and all the authors contributed to the discussion about the conclusion. All authors have read and
agreed to the published version of the manuscript.
Funding: This research was funded by the National Natural Science Foundation of China, grant
number 52074079, and Fundamental Research Funds of the Central Universities of China, grant
number N2125018.
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: Data is contained within the article.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Visuri, V.-V.; Vuolio, T.; Haas, T.; Fabritius, T. A Review of Modeling Hot Metal Desulfurization. Steel Res. Int. 2020, 91,
1–25. [CrossRef]
2. Schrama, F.N.H.; Beunder, E.M.; Berg, B.V.D.; Yang, Y.; Boom, R. Sulphur removal in ironmaking and oxygen steelmaking.
Ironmak. Steelmak. 2017, 44, 333–343. [CrossRef]
3. Kanbara, K.; Nisugi, S.; Shiraishi, O.; Katakeyama, T. Desulfurization approach with mechanically stirred. Tetsu-To-Hagané 1972,
58, S26–S34.
4. Nakai, Y.; Sumi, I.; Kikuchi, N.; Tanaka, K.; Miki, Y. Powder Blasting in Hot Metal Desulfurization by Mechanical Stirring Process.
ISIJ Int. 2017, 57, 1029–1036. [CrossRef]
5. Wang, Q.; Jia, S.; Tan, F.; Li, G.; Ouyang, D.; Zhu, S.; Sun, W.; He, Z. Numerical Study on Desulfurization Behavior during
Kanbara Reactor Hot Metal Treatment. Metall. Mater. Trans. B. 2021, 52, 1085–1094. [CrossRef]
6. Shao, P.; Zhang, T.; Liu, Y.; Zhao, H.; He, J. Numerical Simulation on Fluid Flow in Hot Metal Pretreatment. J. Iron Steel Res. Int.
2011, 18, 129–134.
7. Yamamoto, T.; Kato, W.; Komarov, S.V.; Ishiwata, Y. Investigation on the Surface Vortex Formation during Mechanical Stirring
with an Axial-Flow Impeller Used in an Aluminum Process. Met. Mater. Trans. A 2019, 50, 2547–2556. [CrossRef]
Metals 2021, 11, 1596 18 of 18

8. Yamamoto, T.; Fang, Y.; Komarov, S.V. Surface vortex formation and free surface deformation in an unbaffled vessel stirred by
on-axis and eccentric impellers. Chem. Eng. J. 2019, 367, 25–36. [CrossRef]
9. Kato, K.; Yamamoto, T.; Komarov, S.V.; Taniguchi, R.; Ishiwata, Y. Evaluation of Mass Transfer in an Aluminum Melting Furnace
Stirred Mechanically during Flux Treatment. Mater. Trans. 2019, 60, 2008–2015. [CrossRef]
10. Komarov, S.; Yamamoto, T.; Arai, H. Incorporation of Powder Particles into an Impeller-Stirred Liquid Bath through Vortex
Formation. Materials 2021, 14, 2710. [CrossRef]
11. He, M.; Wang, N.; Hou, Q.; Chen, M.; Yu, H. Coalescence and sedimentation of liquid iron droplets during smelting reduction of
converter slag with mechanical stirring. Powder Technol. 2020, 362, 550–558. [CrossRef]
12. He, M.; Wang, N.; Chen, M.; Li, C. Distribution and motion behavior of desulfurizer particles in hot metal with mechanical
stirring. Powder Technol. 2020, 361, 455–461. [CrossRef]
13. Li, M.; Tan, Y.; Sun, J.; Xie, D.; Liu, Z. Drawdown mechanism of light particles in baffled stirred tank for the KR desulphurization
process. Chin. J. Chem. Eng. 2019, 27, 247–256. [CrossRef]
14. Li, M.; Tan, Y.; Liu, Y.; Sun, J.; Xie, D.; Liu, Z. Effects of geometrical and physical factors on light particles dispersion by agitation
characteristic curve. Chin. J. Chem. Eng. 2019, 27, 2313–2324. [CrossRef]
15. Nakai, Y.; Sumi, I.; Matsuno, H.; Kikuchi, N.; Kishimoto, Y. Effect of Flux Dispersion Behavior on Desulfurization of Hot Metal.
ISIJ Int. 2010, 50, 403–410. [CrossRef]
16. Nakai, Y.; Hino, Y.; Sumi, I.; Kikuchi, N.; Uchida, Y.; Miki, Y. Effect of Flux Addition Method on Hot Metal Desulfurization by
Mechanical Stirring Process. ISIJ Int. 2015, 55, 1398–1407. [CrossRef]
17. Ji, J.-H.; Liang, R.-Q.; He, J.-C. Simulation on Mixing Behavior of Desulfurizer and High-sulfur Hot Metal Based on Variable-
velocity Stirring. ISIJ Int. 2016, 56, 794–802. [CrossRef]
18. Sahu, M.K.; Sahu, R.K. Optimization of Stirring Parameters Using CFD Simulations for HAMCs Synthesis by Stir Casting Process.
Trans. Indian Inst. Met. 2017, 70, 2563–2570. [CrossRef]
19. Liu, Y.; Zhang, T.-A.; Sano, M.; Wang, Q.; Ren, X.-D.; He, J.-C. Mechanical stirring for highly efficient gas injection refining. Trans.
Nonferrous Met. Soc. China 2011, 21, 1896–1904. [CrossRef]
20. Li, Q.; Shen, X.; Guo, S.; Li, M.; Zou, Z. Computational Investigation on Effect of Impeller Dimension on Fluid Flow and Interface
Behavior for Kanbara Reactor Hot Metal Treatment. Steel Res. Int. 2021, 92, 1–15. [CrossRef]
21. Li, M.; Shao, L.; Li, Q.; Zou, Z. A Numerical Study on Blowing Characteristics of a Dynamic Free Oxygen Lance Converter for
Hot Metal Dephosphorization Technology Using a Coupled VOF-SMM Method. Met. Mater. Trans. A 2021, 52, 1–12. [CrossRef]
22. Ubbink, O.; Issa, R. A Method for Capturing Sharp Fluid Interfaces on Arbitrary Meshes. J. Comput. Phys. 1999, 153,
26–50. [CrossRef]
23. Hirt, C.W.; Nichols, B.D. Volume of fluid (VOF) method for the dynamics of free boundaries. J. Comput. Phys. 1981, 39,
201–225. [CrossRef]
24. Chuprov, P.; Utkin, P.; Fortova, S. Numerical Simulation of a High-Speed Impact of Metal Plates Using a Three-Fluid Model.
Metals 2021, 11, 1233. [CrossRef]
25. Wang, Y.; Cao, L.; Cheng, Z.; Blanpain, B.; Guo, M. Mathematical Methodology and Metallurgical Application of Turbulence
Modelling: A Review. Metals 2021, 11, 1297. [CrossRef]
26. Luo, J.Y.; Gosman, A.D.; Issa, R.I.; Middleton, J.C.; Fitzgerald, M.K. Full flow field computation of mixing in baffled stirred
vessels. Chem. Eng. Res. Des. 1993, 71, 342–344.
27. Torotwa, I.; Ji, C. A Study of the Mixing Performance of Different Impeller Designs in Stirred Vessels Using Computational Fluid
Dynamics. Designs 2018, 2, 10. [CrossRef]
28. Tanaka, R.; Uddin, A.; Kato, Y. Flow Characteristics Related to Liquid/liquid Mixing Pattern in an Impeller-stirred Vessel. ISIJ Int.
2018, 58, 620–626. [CrossRef]
29. Wang, Q.; Liu, Y.; Cao, Y.; Li, G. Study on Sulfur Transfer Behavior during Refining of Rejected Electrolytic Manganese Metal.
Metals 2019, 9, 751. [CrossRef]

You might also like