You are on page 1of 9

Hydrometallurgy 131-132 (2013) 67–75

Contents lists available at SciVerse ScienceDirect

Hydrometallurgy
journal homepage: www.elsevier.com/locate/hydromet

Numerical modelling of hydrodynamics and gas dispersion in an autoclave


H. Appa a, b, c,⁎, D.A. Deglon a, b, c, C.J. Meyer c, d
a
Minerals to Metals Initiative (MtM), Department of Chemical Engineering, University of Cape Town, Private Bag Rondebosch, Cape Town 7700, South Africa
b
Centre for Minerals Research, Department of Chemical Engineering, University of Cape Town, Private Bag Rondebosch, Cape Town 7700, South Africa
c
Centre for Research in Computational and Applied Mechanics, University of Cape Town, Private Bag Rondebosch, Cape Town 7700, South Africa
d
Department of Mechanical and Mechatronic Engineering, Faculty of Engineering, University of Stellenbosch, Private Bag X1, Matieland, 7602, Stellenbosch, South Africa

a r t i c l e i n f o a b s t r a c t

Article history: This paper investigates the numerical modelling of hydrodynamics and gas dispersion in the first compart-
Received 1 February 2012 ment of an autoclave. A scaled down model of an 80 m 3 Sherritt Gordon horizontal autoclave agitated by a
Received in revised form 24 September 2012 radial Smith turbine was used. The CFD simulations were implemented using the commercial code, ANSYS
Accepted 6 October 2012
Fluent 13. The CFD model was validated using experimentally determined power draw, velocity flow field
Available online 12 October 2012
and bubble size data. Hydrodynamics and gas dispersion displayed unusual characteristics compared to
Keywords:
conventional stirred tanks. Fluid flow leaving the impeller was found to have a negative inclination to the
Autoclave horizontal. Mixing was found to be asymmetric and non-ideal. Gas dispersion was found to be relatively
Computational fluid dynamics poor, non-homogeneous with accumulation of oxygen gas below the impeller, asymmetric and with low
Modelling gas hold-ups.
Multiphase © 2012 Elsevier B.V. All rights reserved.

1. Introduction flow patterns in stirred tanks. The double-loop structure changes to


a single-loop when the impeller distance from the impeller and the
Autoclaves are commonly used for leaching of minerals (Nicolle bottom of the vessel is reduced (Montante et al., 2001; Nienow,
et al., 2009; Roux et al., 2009). There have been numerous studies 1968). Turbulent kinetic energy and velocity in the region close to
on leaching and leach chemistry in autoclaves (Georgiou and the impeller has been observed to peak close to the impeller and dis-
Papangelakis, 2004; Lamya, 2007). However, few studies have fo- sipate in the regions away from the impeller.
cused on hydrodynamics and gas dispersion in autoclaves and even Gas dispersion is controlled by hydrodynamics. The general flow
fewer have developed computational models to predict these flow field in a stirred tank with a radial impeller is a non-homogeneous
phenomena. This lack of understanding of hydrodynamics and gas gas distribution, with high gas accumulation in the recirculation
dispersion is detrimental to the efficiency of the leaching process loops and close to the impeller tip due to the low-pressure in those
since they affect gas–liquid mass transfer. The latter is a critical regions. The spatial bubble size distribution is non-homogeneous, as
component in leach kinetics and may be the rate limiting step a result of the complex interaction between the local gas-hold up,
(Crundwell, 1995) especially when the reaction is fast (Jakobsen, 2008). local bubble size and local turbulence intensity (Bakker and Van
Hydrodynamics in stirred tanks have been extensively studied Den Akker, 1994). Larger bubbles are formed in regions of higher
both experimentally and numerically. In both cases the research gas hold-up due to the higher probability of collision. Smaller bubbles
studies have placed emphasis on flow patterns, velocity flow field, are formed due to break-up which is controlled by the turbulent ki-
turbulent dissipation rate and kinetic energy. These are important to netic energy within the eddies with wavelengths of the order of the
understand the mixing characteristics of stirred reactors. The flow bubble diameter. Therefore, smaller bubbles are expected closer to
pattern in a baffled cylindrical vessel with a radial impeller, such the impeller tip. Moreover, the increase in the number of bubbles
as the Rushton turbine is well known, with two recirculation loops; due to break-up also leads to an increase in coalescence due to the in-
one below and one above the impeller. These are caused by the crease in bubble collision rate. Thus, bubble size will depend on the
high radial jet which splits into two upon hitting the vessel wall. In dominating mechanism.
the absence of baffles a vortex is formed on the free surface of the In the last few decades efforts have been steered towards the
liquid (Rushton et al., 1994). A similar observation is made if the im- development of predictive methods using Computational Fluid Dy-
peller shaft is not in the centre of the vessel, with two vortical struc- namics (CFD). Various CFD methodologies have been presented
tures; one above and one below the impeller (Galletti et al., 2009). to simulate fluid flow in stirred vessels for both single phase (Aubin
The impeller clearance from the bottom of the vessel also affects the et al., 2004; Brucato et al., 1998; Deglon and Meyer, 2006; Javed et
al., 2006; Ranade et al., 1989) and gas–liquid systems (Bakker and
⁎ Corresponding author. Tel.: +27 83 315 54 08. Van Den Akker, 1994; Gimbun et al., 2009; Kerdouss et al., 2008;
E-mail address: HarishAppa@gmail.com (H. Appa). Montante et al., 2008; Ranade and Deshpande, 1999). However,

0304-386X/$ – see front matter © 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.hydromet.2012.10.006
68 H. Appa et al. / Hydrometallurgy 131-132 (2013) 67–75

few studies have been focused on the CFD modelling of autoclaves 2.2. Power
(Nicolle et al., 2009).
The purpose of this research is to develop a CFD model to predict Power input was determined from the torque experienced by the
hydrodynamics and gas dispersion in the first compartment of an motor. The motor was mounted between bearings to allow free rota-
autoclave and to validate the model against experimental data. Fur- tion and any torque experienced due to the inertia on the impeller
thermore, the numerical results and experimental data will be used blades resulted in a rotation. The resultant moment on a lever arm
to illustrate key features of hydrodynamics and gas dispersion in fitted to the motor was measured by a load cell. The power was deter-
autoclaves. mined from the product of the torque and the impeller speed.

2.3. Velocity flow field


2. Experimental
Velocity flow field was measured using a Dantec Dynamics 2-D
2.1. Autoclave PIV system. The autoclave was seeded with Polyamid Seeding Parti-
cles (PSP) with a diameter of 50 μm. A Nd-YAG double pulsed laser
A scaled down model of the first compartment of an 80 m 3 beam was used to create a 1 mm thick light sheet. The measurements
Sherritt Gordon horizontal autoclave was used for this study. The were carried out at the centre of the tank. A CCD camera, with a res-
first compartment was geometrically scaled down by a factor of 7 olution of 1024 × 1028 pixels, synchronised with the laser pulse was
and had a capacity of 60 l. The latter was chosen so that the vessel placed at a right angle to the light sheet. The occurrence of a small
was large enough to have similar hydrodynamics characteristics to surface vortex which allowed air to be drawn into the impeller was
industrial autoclaves, yet sufficiently small to conduct laboratory found at about 225 rpm. These could have led to inaccuracies in the
characterisation experiments for CFD validation. The vessel has two PIV measurements. Therefore, three different impeller speeds were
baffles, one on each side of the vessel as shown in Fig. 1. The system investigated; 150, 200 and 250 rpm.
was agitated with a six bladed Smith turbine impeller placed at a The exposure time delay was determined after carrying out pre-
distance of 102 mm above the bottom surface. The dimensions in liminary experiments for each impeller speed and at different planes.
mm are shown in Fig. 2. A diamond shaped sparger, placed below Based on the results the exposure time delay was set to 1000 μs. The
the impeller at a distance of 48 mm below the impeller, was used. DANTEC software was used to calculate the velocity vectors. The anal-
Oxygen was sparged through four equidistant 2 mm holes drilled on ysis was performed initially by dividing the captured images into
the four sides of the sparger. interrogation areas of 32× 32 pixels. The average flow field was obtained
The experimental programme is summarised in terms of experi- by time averaging the data over the 150 images, which were found
mental conditions and experimental variables in Table 1. The vessel sufficient to provide reasonable results.
was filled with 60 l of water and operated at ambient temperature,
25 °C. The maximum impeller speed of 395 rpm was determined by 2.4. Bubble size distribution
scaling on a constant power per unit volume basis. Oxygen gas was
used for the two phase experiments. The oxygen flow rate of 4.37 l An Anglo American Platinum Bubble Sizer (AMPBS), which is
min −1 was determined by scaling on a constant flow rate per unit based on a photographic technique, was used to determine the bub-
volume basis. This resulted in a substantially lower superficial gas ble size distribution. The system was equipped with a 12.0 mega
velocity than that in the industrial autoclave. However, scaling on pixel camera. Here, bubbles rise up a tube until they reach a viewing
a constant superficial gas velocity would have resulted in very chamber where images are taken. The bubble size was extracted
high flow rates, well outside the operating window of the laboratory using the AMPBS software application. In order to get a reasonable
vessel. bubble size distribution it was decided to take samples at 2 different
positions; firstly at a point in the impeller stream (Sample 1) and sec-
ondly in the bulk region (Sample 2). The positions where Samples 1
and 2 were taken were at a distance of 120 mm and 205 mm from
the bottom of the vessel respectively. After the system was allowed
to stabilise a total of 100 photos were taken at intervals of 5 s. The
number of bubbles determined for each experimental run depended
on the operating conditions and had an average of 600 images mea-
sured for every operating condition investigated.
Gas hold-up values are normally used for the characterisation of
gas dispersion and consequently for validation of CFD models. There
are different techniques for the measurements of gas hold-up in
mixing vessels. Two common methods are, measuring the change in
liquid volume due the gas phase or measuring density changes using
pressure probes. However, the gas hold-up was considered to be too
low to be accurately measured using any of these methods.

3. Computational model

The CFD simulations were implemented using the commercial


CFD code, ANSYS Fluent 13.

3.1. Single phase: water

3.1.1. Geometry and boundary conditions


The geometry of the vessel was based on the experimental rig,
Fig. 1. Pilot-scale experimental rig of first compartment of autoclave. with the same dimensions. Fig. 3 shows the geometry used for the
H. Appa et al. / Hydrometallurgy 131-132 (2013) 67–75 69

Fig. 2. Tank and impeller dimensions.

CFD model. Due to the asymmetry of the vessel the fluid flow in exist between the impeller and the curved sides of the vessel as
the system was expected to be asymmetrical. Therefore, a full well as the baffles. The sliding mesh was run for 20 impeller rotations
3-dimensional geometry was simulated. The impeller and the sparger for all single phase simulations. In order to implement the MRF or the
were modelled as infinitely thin walls and the shaft was extended to SM model an impeller zone, a cylindrical region, around the impeller
the bottom of the tank to facilitate the meshing of the vessel. Standard has to be defined. According to Lee and Yianneskis (1994) this region
wall functions as proposed by Launder and Spalding (1974) were has to be of a minimum size to be able to capture the strongly period-
used to account for the viscous flow near the solid surfaces. The free sur- ic flows in a stirred system. However, due to the geometry of the ves-
face was modelled with a symmetry boundary. Two different meshes sel and the position of the sparger the dimensions suggested by Lee
were used for this study, 322,749 cells (Grid 1) and 1,061,360 cells and Yianneskis (1994) were altered. The cylindrical region used to
(Grid 2). Grid 1 had 90,054 and 232,695 cells for the impeller and describe the impeller region had a height that was 2.5 times the
bulk regions respectively. Grid 2 had 114,752 and 946,608 cells for the blade height and a width 1.5 times the impeller diameter.
impeller and bulk regions respectively.

3.1.3. Turbulence model


3.1.2. Impeller rotation model The choice of the turbulence model also affects the flow predic-
The flow in mixing vessels is quite complex. For instance, in stirred tion. The turbulence model does not highly influence the prediction
tanks the baffle-impeller interaction results in periodic and time de- of mean velocity but poor prediction of turbulent kinetic energy
pendent flows. There are various models that have been used to nu- is well known when using the k − ε model. However, the research
merically account for these effects, namely the Snapshot approach presented by Deglon and Meyer (2006) showed that the discrepan-
(Khopkar et al., 2005; Ranade and Deshpande, 1999; Ranade and cies were mainly due to numerical errors rather than a deficiency of
Van Den Akker, 1994), the Multiple Reference Frames (MRF) model the model. Similar findings were made by Wechsler et al. (1999) and
and the Sliding Mesh (SM) method have been used in either Aubin et al. (2004). The variances can be resolved by using a fine
multiphase or single phase systems. The system was simulated with mesh and a higher order discretisation scheme. Therefore, the standard
both the MRF and the Sliding mesh models. The MRF model was
used to start up the simulation in order to reduce computational ex-
pense. The SM model was used to solve the flow field completely
and also to account for any time dependent interaction that could

Table 1
Experimental conditions and variables.

Experimental Conditions

Tank volume 60 l
Medium Water
Gas Oxygen
Gas flow rate 4.37 l min−1
Temperature 25 °C

Experimental Variables

Impeller Speed 150, 200, 225, 250


300, 350, 395 rpm Fig. 3. Geometrical representation of numerical model.
70 H. Appa et al. / Hydrometallurgy 131-132 (2013) 67–75

k − ε model was used to model the turbulent flow regime for the 3.2.3. Discretisation scheme and pressure–velocity coupling
autoclave. Momentum and volume fraction equations were discretised
using the QUICK scheme. First order schemes were used for the
3.1.4. Discretisation scheme and pressure–velocity coupling discretisation of the turbulent kinetic energy and dissipation rate.
The order of the discretisation schemes when modelling stirred The phase-coupled SIMPLE algorithm was used for the pressure ve-
tanks has been reported to highly influence the prediction of the locity coupling.
turbulent kinetic energy (Deglon and Meyer, 2006). However, if the
grid is not too coarse the mean velocity is negligibly affected by the 4. Autoclave CFD model validation
discretisation schemes. Similar findings were made by Aubin et al.
(2004) while investigating effects of various discretisation schemes. 4.1. CFD model
Therefore, the first order upwind differencing scheme was used for
the start of the simulations to discretise the momentum and turbu- Single phase and multiphase simulations were done using an
lence equations. Then the higher order QUICK scheme was used for AMD Quad Core 2.15 GHz CPU with 8 GB of RAM and an AMD dual
the discretisation of the equations. The SIMPLE algorithm was used Quad Core 2.2 GHz CPU with 16 GB of RAM respectively. For both
for the pressure–velocity coupling. multiphase and single phase systems all cores were used, that is 8 and
4 respectively. The simulations were considered to have converged
3.2. Multiphase: water–oxygen when the scaled residuals were below 1×10−3. The average running
time for the multiphase simulations was about three weeks. The CFD
The gas–liquid system, with water and oxygen gas, was modelled models were validated against experimental power draw, velocity flow
with the commonly used Euler-Euler model (Aubin et al., 2004; field and Sauter mean bubble diameter. As indicated previously, gas
Gimbun et al., 2009; Kerdouss et al., 2008; Khopkar et al., 2005; hold-up was too low to be measured accurately and therefore could not
Morud and Hjertager, 1996; Ranade and Van Den Akker, 1994; be used for CFD validations. For the bubble size, even though as reported
Scargiali et al., 2007). The dispersed k − ε turbulence model as simi- by Laakkonen et al. (2007) a large number of classes are required to sim-
larly used by Kerdouss et al. (2008) was employed to simulate the ulate satisfactory gas liquid systems, the discrete model with 11 classes
turbulent flow. When using the k − ε model the primary phase controls accurately predicted the bubble size. Similarly, Kerdouss et al. (2008)
the motion of the secondary phase and the influence of the bubbles on found that 13 classes gave satisfactory predictions for simulating a gas–
the flow field is negligible. Therefore, to ensure that the effect of the liquid system for stirred tanks. This shows that the number of classes
secondary phase is minimised and to avoid any numerical instability used is problem dependent and a prior knowledge of the bubble size
the bubble size should be smaller than the cell size. To this end, and distribution helps in choosing the optimal number of classes.
based on the results of the single phase simulations, Grid 1 was used
to model the gas–liquid system. Turbulent dispersion is taken into ac- 4.2. Hydrodynamics
count by the dispersed k − ε turbulence model (Simonin and Viollet,
1990). The Schiller–Naumann model (Schiller and Naumann, 1933) 4.2.1. Power
was used to account for the interphase drag. The result for the power draw comparing the experimental and
numerical data is plotted in Fig. 4. Power from the CFD simulations
3.2.1. Boundary conditions was calculated from the torque experienced by the impeller and the
The boundary conditions used are similar to the ones described known impeller speed. The torque was determined from the sum of
for the single phase system. However, to account for the gas phase, moments (sum of cross product of the pressure and viscous forces
the sparger was modelled as a velocity inlet boundary and the outlet with the moment vector) on both sides of the impeller blades. The
was modelled as a degassing boundary, where a velocity boundary numerical data for the two cases for Grid 1 and Grid 2 are also com-
was used. The velocity of the gas at the sparger was based on the vol- pared. Power as expected increases as impeller speed is increased.
umetric flow rate. For the degassing boundary the velocity and the There is also a good correlation between the experimental and nu-
volume fraction of the gas at the boundary was set to the value at merical data. However, a small over-prediction is observed at 350
one cell below the boundary. A no-slip wall was used for the liquid and 395 rpm. This could be a consequence of not considering the
phase, with water only allowed to move along the boundary and
with the normal component of the velocity set to zero.
0.16
Experimental
3.2.2. Gas phase Numerical, Grid 1
Numerical, Grid 2
The dispersed phase was modelled using a population balance 0.14
model. The bubble size distribution was simulated using the discrete
method. Using this method the particle size distributions are repre- 0.12
sented in terms of classes or bins. It allows for the direct simulation
of the bubble size distribution but at the expense of extra computa- 0.1
Power (kW)

tional cost due to the large number of classes required (Laakkonen


et al., 2007). However, for this study with the availability of experi- 0.08
mental data 11 classes were sufficient. The bubble size distribution,
which was used as input for the model, was divided into the classes 0.06
according to the following equation:
0.04
viþ1 q
¼2 ð1Þ 0.02
vi

0
where v is the volume of the bubble and q is the ratio factor that can 150 200 250 300 350 400
be obtained using a geometric progression as well as the minimum Impeller speed (rpm)
and maximum bubble volumes. The Luo models (Luo and Svendsen,
1996) were used to simulate bubble breakage and coalescence. Fig. 4. Power — CFD vs Experimental.
H. Appa et al. / Hydrometallurgy 131-132 (2013) 67–75 71

effect of the free surface in the numerical model. At impeller speeds of


350 and 395 rpm the liquid was more susceptible to the occurrence of
a free vortex and there was a higher degree of recirculation at the sur-
face of the tank. The effect of the free surface was not implemented as
it was found that the computational effort required would not have
been worth it. Moreover, it can be observed that there is no difference
in power input for the two grids investigated. This suggests that Grid
1 is fine enough to predict power draw and also that the discretisation
schemes used had a negligible effect on the solution. Therefore, Grid 1
with the QUICK scheme is appropriate for the system.

4.2.2. Velocity flow field


The velocity flow field on a vertical plane in the center of the tank
at an impeller speed of 150 rpm is shown in Figs. 5–7. Fig. 5 depicts
the time-averaged velocity flow field on the yz-plane from the PIV
measurements. Figs. 6 and 7 show the instantaneous velocity flow
field from the CFD simulations, with the impeller at two different po-
sitions, Position 1 and Position 2 respectively. Position 1 is with the
vertical plane behind one of the impeller blades and Position 2 with Fig. 6. Velocity flow field at 150 rpm at Position 1 — CFD.
the vertical place midway between two of the impeller blades. The
typical flow pattern that is commonly seen in a flat bottom stirred
tank with a radial impeller is observed. The double-loop structure is
noticeable, with a recirculation loop below and above the impeller. The discrepancies could be due to an inadequacy of the CFD model,
However, the horizontal radial jet as commonly observed with radial resulting in an under-prediction of the velocity in the impeller region.
impellers is not reproduced. It can be seen from Figs. 5–7 that both It is also known to be more difficult to obtain good velocity predictions
the CFD prediction and the PIV measurement show that the fluid in this region. In addition, no significant difference could be observed
leaves the impeller at an angle to the horizontal. It is suspected that between the prediction when using Grid 1 and Grid 2 respectively.
the flow is either being forced down by the axial component of the
fluid from the bulk flow entering the impeller stream or ‘attracted’
(Montante et al., 2001) to the bottom of the vessel. This inflection 4.3. Gas dispersion
can also be either due to the effect of the low pressure in the lower
recirculation loop or an ’attraction’ due to the ‘Coanda’ effect as a result The numerical prediction and experimental values of the Sauter
of the bottom surface of the tank (Montante et al., 2001). mean diameter versus the impeller speed is shown in Fig. 11. It can
A flow visualisation study, where the flow pattern was highlighted be seen that the experimental and predicted Sauter mean diameter
with a dye, was used to visually confirm this unusual finding. Fig. 8 are comparable over the range of impeller speeds investigated, with
shows the direction of the fluid in the impeller discharge region at an average of 4% and a maximum error of 12%. The result shows
150 rpm at the centre of the vessel on one side of the vessel. This fig- that the discrete model accurately predicts the bubble size distribu-
ure clearly shows the inclined radial jet leaving the impeller. tion with 11 classes. This was possible because of the small variation
Figs. 9 and 10 show the radial velocity along a line at 10 mm away in the bubble size, which was between 1 mm and 6.6 mm. Further-
from the impeller tip at 150 and 200 rpm respectively. It can be ob- more, an increase in bubble diameter is observed while increasing im-
served that the velocity peaks below the impeller, which is at peller speed. This is unusual as the bubble size is expected to decrease
102 mm from the bottom surface of the vessel. This confirms the with impeller speed. However, for the autoclave at these speeds inter-
downward inclination of the fluid flow. Moreover, good predictions nal recirculation was found to increase, leading to a higher probability
were obtained for the velocity profiles in the bulk region. However, of collision and coalescence, to consequently result in an increase in
a disagreement in the velocities can be seen in the impeller stream. bubble size.

Fig. 5. Velocity flow field at 150 rpm — PIV. Fig. 7. Velocity flow field at 150 rpm at Position 2 — CFD.
72 H. Appa et al. / Hydrometallurgy 131-132 (2013) 67–75

0.35
Numerical − Grid 1
Numerical − Grid 2
0.3 PIV

0.25

Position, z (m)
0.2

0.15

0.1

0.05

0
−1 −0.5 0 0.5
Velocity in y direction (ms−1)
Fig. 10. Velocity profile in y direction at 200 rpm.

Fig. 8. Velocity flow field at 150 rpm — flow visualisation.

5.1. Hydrodynamics

5.1.1. Power
In summary, results from the single phase simulations showed Power draw was found to increase with increasing impeller speed.
that the standard k − ε model, coupled with the sliding mesh model The Reynolds number was found to vary between 7.34 × 10 4 and
and high order discretisation schemes gave good prediction for the 1.95 × 10 5. The typical turbulent power number, Np, varies between
velocity flow field. These findings are similar to CFD methodologies 2.8 and 3.2 for the Smith turbine. For the impeller speeds investigated
used for stirred tanks. The gas–liquid system was simulated with the the power number was found to vary between 4.1 and 5.6, with the
Euler-Euler model coupled with a population balance model. Using lower power numbers at the higher impeller speeds. These power
this approach and the discrete solution method with 11 classes gave rea- numbers are higher than the typical literature values as reported for
sonably good predictions of the Sauter mean diameter. stirred tanks.

5.1.2. Velocity flow field


5. Autoclave hydrodynamics and gas dispersion The experimental time-averaged velocity flow fields at 150, 200
and 250 rpm on a vertical plane in the middle of the tank are shown
The primary objective of this paper was to develop and validate a in Figs. 12, 13 and 14 respectively. It can be seen that as the impeller
CFD model in order to simulate hydrodynamics and gas dispersion in speed increases the angle of the radial flow to the horizontal de-
the first compartment of an autoclave. Based on the CFD models and creases. Moreover, it can be observed that the double loop structure is
experimental data, key hydrodynamics and gas dispersion features transitioning to a single loop structure. This phenomenon is normally ob-
are described in the following sections. served in stirred tanks with radial impellers when the distance between

0.35 3.5
Numerical − Grid 1
Numerical − Grid 2
0.3 PIV

3
0.25
Position, z (m)

2.5
d32 (mm)

0.2

0.15
2

0.1
1.5
0.05 Impeller Experimental
Impeller Numerical
Vessel Experimental
Vessel Numerical
0 1
−0.8 −0.6 −0.4 −0.2 0 0.2 0.4 0.6 200 220 240 260 280 300 320 340 360 380 400
Velocity in y direction (ms−1) Impeller speed (rpm)
Fig. 9. Velocity profile in y direction at 150 rpm. Fig. 11. Sauter mean diameter — CFD and experimental.
H. Appa et al. / Hydrometallurgy 131-132 (2013) 67–75 73

Fig. 14. Velocity flow field at 250 rpm — PIV.


Fig. 12. Velocity flow field at 150 rpm — PIV.

the impeller and the bottom of the vessel is reduced (Montante et al., the flow is dependent on the impeller blade position. The asymmetry
2001; Nienow, 1968). Therefore, it can be inferred that the flow pattern of the flow field is due to the geometry of the vessel and also the
is dependent on the impeller speed. The downward inclination of the effect of the shape of the impeller blade on the flow pattern. It is
radial jet is also observed in dished bottom vessel as reported by Deen also observed that the lower recirculation loop disappears on the
et al. (2002). Therefore, the downward inflection can be due to a contri- right hand side of the vessel.
bution of the impeller clearance as well as the geometry of the vessel. The downward inflection is also observed at both Position 1 and
The effect of the curved bottom on the recirculation loop below the im- Position 2. The angle of the radial flow to the horizontal is more pro-
peller increases with impeller speed, resulting in a more or less single nounced than at 150 rpm. This is believed to be a result of a predom-
loop structure. inant axial flow at this speed such that the flow leaving the impeller is
Figs. 15 and 16 show the instantaneous predicted velocity vectors immediately entrained by the axial flow. In addition, the effect of the
on a vertical plane in the centre of the tank at 395 rpm at two differ- curved blades becomes more substantial at 395 rpm resulting in the
ent impeller positions. The impeller blade is on the plane at Position 1 fluid in the impeller discharge to have a greater angle of inflection.
and the plane is midway between two impeller blades. There are two
main observations; firstly the flow field is asymmetric and secondly
5.2. Gas dispersion

5.2.1. Bubble size


Bubble size in terms of local Sauter mean diameter is shown in
Fig. 17. It can be seen that the bubbles below the impeller region are
larger than 2 mm, the size of bubbles at the sparger. However, in the
impeller region a decrease is observed, mainly as a result of high break-
age rates caused by a higher turbulent kinetic energy in that region. In
the bulk region coalescence is observed. The higher coalescence rate
in that region is due to a higher probability of collision between bubbles.
Furthermore, when looking at the vessel globally, with most bubbles
greater than 2 mm in diameter, it can be deduced that the system
is mainly coalescence controlled and with small breakage in the impel-
ler region.
The variation of bubble size in the autoclave is different to the one
generally observed in stirred tanks. Gas–liquid mass transfer is better
when the inter-facial area is high which occurs in systems where
bubble size is small. With the current system which is mainly con-
trolled by coalescence and with relatively large bubbles, mass transfer
is expected to be adversely affected.

5.2.2. Gas hold-up


Fig. 18 shows the volume fraction of oxygen on a vertical plane in
the centre of the tank. The CFD simulation was carried out at 395 rpm
with a gas flow-rate of 4.37 l min −1. The gas hold-up values are low.
Fig. 13. Velocity flow field at 200 rpm — PIV. This is due to a combination of the low superficial gas velocity used,
74 H. Appa et al. / Hydrometallurgy 131-132 (2013) 67–75

Fig. 15. Velocity flow field on vertical plane across tank at Position 1 — CFD.

Fig. 16. Velocity flow field on vertical plane across tank at Position 2 — CFD.

the large bubbles and poor dispersion of gas. The distribution of gas is 6. Conclusions
distinctly non-homogeneous and asymmetrical. Most of the gas is
trapped under the impeller and in the two large recirculating loops. The purpose of this study was to develop and validate a CFD model
There are entire regions of the vessel where there is practically no of an autoclave. The study was conducted using a scaled-down, 60 l
oxygen. model of the first compartment of an industrial autoclave.
Gas dispersion is unlikely to improve at higher superficial gas ve-
locities. Here, the overall gas hold-up would increase but the quality • CFD model: It was found that the same CFD methodology used
of gas dispersion may even deteriorate. The relatively poor gas dis- for stirred tanks was appropriate for the autoclave. For the single
persion would have an impact on mass transfer in the autoclave, phase system, the standard k − ε model with high order QUICK
with no mass transfer occurring in certain regions. scheme gave reasonable predictions of the power draw and velocity

5.53e-03 5.00e-02

5.07e-03 4.50e-02

4.62e-03 4.00e-02

4.17e-03 3.50e-02

3.72e-03 3.00e-02

3.26e-03 2.50e-02

2.81e-03 2.00e-02

2.36e-03 1.50e-02

1.91e-03 1.00e-02

1.45e-03 5.00e-02

1.00e-03 0.00e+02

Fig. 17. Local Sauter mean diameter on vertical plane at 395 rpm. Fig. 18. Volume fraction of oxygen on a vertical plane at 395 rpm.
H. Appa et al. / Hydrometallurgy 131-132 (2013) 67–75 75

flow field. For the two phase system, the Euler-Euler model with Gimbun, J., Rielly, C., Nagy, Z., 2009. Modelling of mass transfer in gas–liquid stirred
tanks agitated by Rushton turbine and CD-6 impeller: a scale-up study. Chem.
the dispersed k − ε model, coupled with the population balance Eng. Res. Des. 87, 437–451.
model, predicted the Sauter mean diameter fairly well. Jakobsen, H., 2008. Chemical Reactor Modeling: Multiphase Reactive Flows. Springer.
• Hydrodynamics: The power number for the Smith turbine was Javed, K., Mahmud, T., Zhu, J., 2006. Numerical simulation of turbulent batch mixing in
a vessel agitated by a Rushton turbine. Chem. Eng. Process. 45, 99–112.
found to be slightly higher than expected. The flow pattern as generally Kerdouss, F., Bannari, A., Proulx, P., Bannari, R., Skrga, M., Labreque, Y., 2008. Two-phase mass
observed in stirred tanks with radial impellers was not reproduced. The transfer coefficient prediction in stirred vessel with a CFD model. Comput. Chem. Eng.
flow in the impeller discharge region was at an angle to the horizontal 32, 1943–1955.
Khopkar, A., Ramnohan, A., Ranade, V., Dudukovic, M., 2005. Gas–liquid flow generated
and the typical double-loop structure changed to a single-loop structure by a Rushton turbine in stirred vessel:CARPT/CT measurements and CFD simula-
with increasing impeller speed. tions. Chem. Eng. Sci. 60, 2215–2229.
• Gas dispersion: Bubble size was found to be dominated by coalescence Laakkonen, M., Alopaeus, V., Aittamaa, J., 2007. Modelling local bubble size distribu-
tions in agitated vessels. Chem. Eng. Sci. 62, 721–740.
with relatively large bubbles (>2 mm) in most regions of the auto-
Lamya, R., 2007. A fundamental evaluation of the atmospheric pre-leaching section of
clave. The gas distribution was distinctly non-homogeneous and asym- the nickel-copper matte treatment process. Ph.D. thesis, Universtiy of Stellenbosch.
metrical, with entire regions of the vessel without oxygen. Launder, B., Spalding, D., 1974. The numerical computation of turbulent flows. Comput.
• Implication for mass transfer: These findings suggest that mass transfer Methods Appl. Mech. Eng. 3, 269–289.
Lee, K., Yianneskis, M., 1994. The extent of periodicity of the flow in vessels stirred by
in autoclaves may be lower than conventional stirred tanks due to Rushton impellers. Am. Inst. Chem. Eng. Ser. 90, 5–18.
non-ideal mixing and relatively poor gas dispersion. The CFD model Luo, H., Svendsen, H., 1996. Theoretical model for drop and bubble breakup in turbu-
presented may be used to investigate phenomena such as the effect lent dispersions. Am. Inst. Chem. Eng. 1 (5), 1225–1233.
Montante, G., Lee, K., Brucato, A., Yianneskis, M., 2001. Experiments and predictions
of scale-up and the prediction of mass transfer in industrial autoclaves. of the transitions of the flow pattern with impeller clearance in stirred tanks.
Comput. Chem. Eng. 25, 729–735.
Montante, G., Horn, D., Paglianti, A., 2008. Gas–liquid flow and bubble size distribution
in stirred tanks. Chem. Eng. Sci. 63, 2107–2118.
Acknowledgements Morud, K., Hjertager, B., 1996. LDA measurements and CFD modeling of gas–liquid flow
in stirred tanks. Chem. Eng. Sci. 51 (1996), 233–249.
Nicolle, M., Nel, G., Plikas, T., Shah, U., Zunti, L., Bellino, M., Pieterse, H., 2009. Mixing
The authors would like to thank the Minerals to Metals SARChI system design for the Tati activox® autoclave. J. South. Afr. Inst. Min. Metall. 109,
Chair and Anglo American Platinum for funding this research. 357–364.
Nienow, A., 1968. Suspension of solid particles in turbine agitated baffled vessels.
Chem. Eng. Sci. 23, 1453–1459.
Ranade, V., Deshpande, V., 1999. Gas–liquid flow in stirred reactors: Trailing vortices
References
and gas accumulation behind impeller blades. Chem. Eng. Sci. 54, 2305–2315.
Ranade, V., Van Den Akker, H., 1994. A computational snapshot of gas–liquid flow in
Aubin, J., Fletcher, D., Xuereb, C., 2004. Modelling turbulent flow in stirred tanks with
baffled stirred reactors. Chem. Eng. Sci. 49 (24B), 5175–5192.
CFD: the influence of the modelling approach, turbulence model and numerical
Ranade, V., Joshi, J., Marathe, A., 1989. Flow generated by pitched blade turbines II:
scheme. Exp. Therm. Fluid Sci. 28, 431–445.
Simulation using the k-ε model. Chem. Eng. Commun. 81, 225–248.
Bakker, A., Van Den Akker, H., 1994. A computational model for the gas–liquid flow in
Roux, J., Toit, M.D., Shklaz, D., 2009. Novel redesign of a pressure leach autoclave by a
stirred reactors. Inst. Chem. Eng. 72 (Part A), 594–606.
South African platinum producer. The Southern African Institute of Mining and
Brucato, A., Ciofalo, M., Grisafi, F., Micale, G., 1998. Numerical prediction of flow fields
Metallurgy, Base Metals Conference.
in baffled stirred vessels: A comparison of alternative modelling approaches.
Rushton, J., Costich, E., Everett, H., 1994. Power characteristics of mixing impellers, part
Chem. Eng. Sci. 53 (21), 3653–3684.
1. Chem. Eng. Prog. 46 (8), 395–404.
Crundwell, F., 1995. Progress in the mathematical modelling of leaching reactors.
Scargiali, F., D'Orazio, A., Grisafi, F., Brucato, A., 2007. Modelling and simulation of
Hydrometallurgy 39, 321–335.
gas–liquid hydrodynamics in mechanically stirred tanks. Chem. Eng. Res. Des.
Deen, N., Solberg, T., Hjertager, B., 2002. Flow generated by an aerated Rushton impeller:
85, 637–646.
two-phase PIV experiments and numerical simulations. Can. J. Chem. Eng. 80, 638–652.
Schiller, L., Naumann, A., 1933. A drag coefficient correlation. Z. Ver. Deutsch. Ing. 77,
Deglon, D., Meyer, C., 2006. CFD modelling of stirred tanks: numerical considerations.
318–320.
Miner. Eng. 19, 1059–1068.
Simonin, C., Viollet, P.L., 1990. Predictions of an oxygen droplet pulverization in a com-
Galletti, C., Pintus, S., Brunazzi, E., 2009. Effect of shaft eccentricity and impeller blade
pressible subsonic coflowing hydrogen flow. Numerical Methods for Multiphase
thickness on the vortices features in an unbaffled vessel. Chem. Eng. Res. Des. 87
Flows FED91, pp. 65–82.
(4), 391–400.
Wechsler, K., Breuer, M., Durst, F., 1999. Steady and unsteady computations of turbu-
Georgiou, D., Papangelakis, V., 2004. Characterization of limonitic laterite and solids dur-
lent flows induced by a 4/45° pitched-blade impeller. J. Fluid Eng. 121, 318–329.
ing sulfuric acid pressure leaching using transmission electron microscopy. Miner.
Eng. 17, 461–463.

You might also like