You are on page 1of 12

Chemical Engineering Science 62 (2007) 7184 7195 www.elsevier.

com/locate/ces

CFD simulations of gasliquidsolid stirred reactor: Prediction of critical impeller speed for solid suspension
B.N. Murthy, R.S. Ghadge, J.B. Joshi
Department of Chemical Engineering, Institute of Chemical Technology, University of Mumbai, Matunga, Mumbai 400 019, India Received 19 April 2007; received in revised form 27 June 2007; accepted 9 July 2007 Available online 13 July 2007

Abstract In this work, simulations have been performed for three phase stirred dispersions using computational uid dynamics model (CFD). The effects of tank diameter, impeller diameter, impeller design, impeller location, impeller speed, particle size, solid loading and supercial gas velocity have been investigated over a wide range. The Eulerian multi-uid model has been employed along with the standard k turbulence model to simulate the gasliquid, solidliquid and gasliquidsolid ows in a stirred tank. A multiple reference frame (MRF) approach was used to model the impeller rotation and for this purpose a commercial CFD code, FLUENT 6.2. Prior to the simulation of three phase dispersions, simulations were performed for the two extreme cases of gasliquid and solidliquid dispersions and the predictions have been compared with the experimental velocity and hold-up proles. The three phase CFD predictions have been compared with the experimental data of Chapman et al. [1983. Particlegasliquid mixing in stirred vessels, part III: three phase mixing. Chemical Engineering Research and Design 60, 167181], Rewatkar et al. [1991. Critical impeller speed for solid suspension in mechanical agitated three-phase reactors. 1. Experimental part. Industrial and Engineering Chemistry Research 30, 17701784] and Zhu and Wu [2002. Critical impeller speed for suspending solids in aerated agitation tanks. The Canadian Journal of Chemical Engineering 80, 16] to understand the distribution of solids over a wide range of solid loading (0.3415 wt%), for different impeller designs (Rushton turbine (RT), pitched blade down and upow turbines (PBT45)), solid particle sizes (1201000 m) and for various supercial gas velocities (010 mm/s). It has been observed that the CFD model could well predict the critical impeller speed over these design and operating conditions. 2007 Elsevier Ltd. All rights reserved.
Keywords: Stirred tank; CFD; EulerianEulerian; Three phase ows; Gasliquidsolid

1. Introduction Stirred reactors involving three phases, gas, liquid and solid, are very common in the chemical and allied industries. The solid phase may act as a catalyst or undergo a chemical reaction. Typical applications include catalytic hydrogenation, FischerTropsch synthesis, oxidation of p-xylene to terephthalic acid, production of polymers using suspension polymerization, oxidative leaching of ores and many other economically important reactions. Various examples of industrial importance have been compiled by Nigam and Schumpe (1996). In some of these applications, the reaction occurs between a dissolved gas and a liquid-phase reactant in the presence of a solid catalyst. In some other cases, the liquid is an inert medium
Corresponding author. Tel.: + 91 22 2414 5616; fax: + 91 22 2414 5614.

E-mail address: jbj@udct.org (J.B. Joshi). 0009-2509/$ - see front matter 2007 Elsevier Ltd. All rights reserved. doi:10.1016/j.ces.2007.07.005

and the reaction takes place between the dissolved gas and the solids. The performance of these reactor types depends upon efcient and simultaneous dispersion of gas and suspension of solid particles. The complexity of the ow generated in the system (3D, recirculating and often turbulent) has compelled the researchers, designers and the practicing engineers to resort to empirical approach to tackle the problems associated with the design, scale-up and optimization of three phase stirred reactors. In order to reduce existing state of empiricism, during the past 30 years, an attempt is being made to understand the underlying uid mechanics and its relationship with the design parameters. In particular, the computational uid dynamics (CFD) and the experimental uid dynamics (EFD) have led to better understanding of the detailed hydrodynamics in single phase ow systems. However, in the case of gasliquid and liquidsolid ows, relatively scanty work has been published using both the EFD and CFD techniques. Further, because of

B.N. Murthy et al. / Chemical Engineering Science 62 (2007) 7184 7195

7185

even additional complexities associated with the three phase systems, practically no published information is available on the CFD simulations of three phase systems. In addition, experimental data for the local velocities and the local gas and solid phase hold-ups are also not available which are useful for the validation of CFD models. In the case of a three-phase stirred tank system, the solid suspension process depends upon the quality of gasliquid dispersion in the absence of solids and the quality of solidliquid dispersion in the absence of gas. The objective of this study was to undertake CFD simulations for the prediction of critical impeller speed for the solid suspension. For this purpose, simulations have been performed for gasliquid, liquidsolid and gasliquidsolid dispersions under different design (mainly impeller design and sparger design) and operating conditions (such as solid loading, particle size, supercial gas velocity and the impeller speed). 2. CFD modeling In the present work, an Eulerian multi-uid model has been adopted to describe the ow behavior of each phase, where the gas, liquid and solid phases are all treated as different continua, interpenetrating and interacting with each other everywhere in the computational domain. In FLUENT, the derivation of the conservation equations for mass and momentum for each of the three phases is done by phase weighted Favre-averaging (Viollet and Simonin, 1994) the local instantaneous balances for each of the phases, and then no additional turbulent dispersion term is introduced into the continuity equation. The pressure eld was assumed to be shared by the three phases, in proportion to their volume fraction. The motion of each phase is governed by respective mass and momentum conservation equations. The continuity equation for each phase is j( j( j(
G G )

j(

S S u S )

jt +

+ .(

S S u S u S ) eff ,S ( uS

= S p + .(S
S S g

+ ( uS )T )) (6)

MI,LS ,

where p is the pressure, eff is the effective viscosity, g is the gravitational acceleration, and MI is the interphase transfer force. The volume fractions satisfy the compatibility conditions G + L + S = 1. 2.1. Interphase momentum transfer Interactions between the phases involve various momentum exchange mechanisms such as the drag, the lift, the added mass force, etc. However, the contribution of drag force has been considered while the effect of the other forces has been ignored. It has been reported that the other forces have no considerable effect on both the gasliquid and solidliquid hydrodynamics in stirred tanks (Ljungqvist and Rasmuson, 2001; Montante et al., 2001and Khopkar et al., 2005). The drag force exerted by the dispersed phase on the continuous phase is calculated as MD,LG = MD,LS = 3 CD,LG 4 dB 3 CD,LS 4 dp
L G |uG L S |uS

(7)

uL |(uG uL ),

(8) (9)

uL |(uS uL ),

where CD is the drag coefcient and d is the diameter of a bubble (dB ) or a particle (dp). The drag coefcient exerted by the gas phase on the liquid phase is obtained by the modied Brucato drag model (Khopkar et al., 2006), which is as follows: dp CD,LG CDO = 6.5 106 CDO
3

jt

+ .(

G G uG ) = 0, L L uL ) = 0, S S uS ) = 0,

(1) (2) (3)

(10)

L L ) + .( jt S S )

jt

+ .(

The drag coefcient exerted by the solid phase on the liquid phase is calculated using the drag law proposed by Pinelli et al. (2001), which is as follows: 16 CDO = [0.4 tanh 1 + 0.6]2 . CD,LS dp 2.2. Turbulence closure In the present work the standard k model for single phase ows has been extended for the three phase ows with extra terms that include interphase turbulent momentum transfer (Elgobashi and Abou-arab, 1983) to take into account the effects of turbulence. The modeling of multiphase turbulent ows is much more complex and computationally expensive for three phase ows mainly because the inuence of the dispersed phases on turbulence of the continuous phase. Therefore, in the present three phase CFD turbulence modeling, it has been assumed that the turbulence in multiphase stirred tanks is (11)

where is the density, is the volume fraction, and u is the velocity vector of each phase. The momentum balance equation for each phase is j(
G G u G )

jt + j( jt +

+ .(

G G uG uG ) eff ,G ( uG

= G p + .(G
G G g L L u L )

+ ( uG ) ))
T

MI,LG ,
L L u L u L ) eff ,L ( uL

(4)

+ .(

= L p + .(L
L L g

+ ( uL )T )) (5)

+ MI,LG + MI,LS ,

7186

B.N. Murthy et al. / Chemical Engineering Science 62 (2007) 7184 7195

restricted within the continuous phase. Further, it is apparent from the literature that both the gasliquid and liquidsolid two phase ows have been successfully simulated by the standard k turbulence model with some modications. The turbulence viscosity of the continuous phase is obtained by the k model:
t,L = C L 2 kL L

(12)

The equations of change for the turbulent kinetic energy (k) and the energy dissipation rate for the liquid phase are given by D L L kL = L Dt + L D L L Dt + L
L (PkL L

t,L kL

kL
kL ,

L ) + L L

(13)
L

= L
L

t,L L

kL

(C 1 PkL C

2 L ) + L L

L.

(14)

Here kL and L represent the inuence of the dispersed phase on the continuous phase and the predictions for turbulence quantities for the dispersed phases are obtained using the Tchen theory of dispersion of discrete particles by homogeneous turbulence (Hinze, 1975). The standard values were used for the turbulence parameters: C 1 = 1.44, C 2 = 1.92, C = 0.09, = 1.3. The value of the molecular viscosity of solid k = 1.0, phase is set to be the same as that of water, since its variation does not bring obvious changes to the simulation results. 3. Method of solution Steady state simulations were performed for the different types of impeller, agitation speeds, particle diameter, solid concentration and supercial gas velocity. The details of the reactor geometry and the operating parameters are given in Tables 1 and 2. In this work, all the simulations have been performed using the commercially available CFD software FLUENT 6.2. The set of governing equations are solved by a nite control volume technique, where the entire vessel has been considered for the simulation. A multiple reference frame (MRF) approach has been used for the simulation of

impeller rotation. In this method, computational domain is divided into impeller zone (rotating reference frame) and stationary zone (stationary reference frame). For all the simulations, the boundary of the rotating domain was positioned at r = 0.16 and 0.10 m z 0.24 m. For the case of Zhu and Wu (2002), the rotating domain was positioned at r = 0.12 and 0.08 m z 0.18 m. Tetrahedral elements were used for meshing the geometry and a good quality of mesh was ensured throughout the computational domain using the GAMBIT mesh generation tool. As regards the particular mesh quality, we have been restricted to use tetra mesh elements due to more number of case studies and complex geometry. However, in this study a very high quality of mesh (skewness < 0.7) has been ensured throughout the computational domain. The number of grid elements in all the three directions in both the impeller and the outer zone were systematically increased. When rening the mesh, care was taken to put most additional mesh elements in the regions of high gradient around the blades and the discharge regions. In order to check the sensitivity of the simulation result on the grid size, the grid spacing was reduced by a factor of 2 till the comparison of the two consecutive cases showed that the reduction of the grid size did not generate a noticeable difference in simulation results. Therefore, grid elements of 600,000700,800 have been used in all the studies. Regarding boundary conditions, tank walls, the impeller surfaces and bafes have been treated as no-slip boundary surfaces with standard wall functions. The bubble size distribution in the stirred tank reactor depends on the design and operating parameters. Unfortunately, experimental data of bubble size distribution in the present case is not available in the published literature. Hence, the mean bubble size of 3 mm has been used for all the simulations. At a liquid surface, a small gas zone was added at the free surface of water, a method that has been reported to dampen instabilities (FLUENT 6.2, 2005) and only gas is allowed to escape using pressure outlet boundary condition which means top surface being exposed to atmospheric pressure. It was initially assumed that the particles were uniformly distributed in the liquid. All terms of the governing equations are discretized using the QUICK scheme. The SIMPLE algorithm has been employed for the pressurevelocity coupling. The convergence criterion (sum of normalized residuals) was set at 104 for all the equations. All the simulations have been carried on the 16 node, 32 processor AMD64 cluster with a clock speed of

Table 1 Geometrical details of the gasliquid and liquidsolid system


References Barresi and Baldi (1987) Mishra and Joshi (1991) Aubin et al. (2004) Angst et al. (2004) Reactor geometry T = 0.39 m, H = 1.19T , C = T /3 T = 0.50 m, H = T , C = T /3 T = 0.19 m, H = T , C = T /3 T = 0.20 m, H = T , C = T /3 Impeller 4-PBTD, D = T /3 6-PBTD, D = T /3 6-PBTD, D = T /2 4-PBTD, D = T /3 Glass particles: dp = 327 m
S

Solids details Glass particles: S = 2660 kg/m3 , dp = 0.15, 0.225, 0.45 mm

Operating variables Solid conc. = 0.5 kg/100 kg, N = Nj s N = 210 rpm, ring sparger = 0.116, sparger location = 0.1 m VG = 0.15 l/s N = 300 rpm, ring sparger = 0.8D , sparger location = 0.6C , VG = 0.042 l/s Solid conc. = 1, 2, 3 vol%, N = 1000 rpm

= 2520 kg/m3 ,

B.N. Murthy et al. / Chemical Engineering Science 62 (2007) 7184 7195


Table 2 Summary of experimental details
References Reactor geometry Impeller DT, D/T = 1/3 DT, PBTD45, PBTD45, D/T = 1/3 Rushton turbine, D/T = 0.3.41 Solids details Glass particles: S = dp = 150, 225, 450 m Quartz particles: S = 2520 kg/m3 , dp = 180, 340, 460, 700 m 2660 kg/m3 , Glass particles: dp = 10.425 m
S

7187

Sparger details Ring type (SR) Pipe (SP) and ring type (SR) Pipe (SP), ring type (SR)

Operating variables Solid conc. = 3 wt%, N = 8.12 rps, VG = 5.20 mm/s Solid conc. = 3.4, 6.6, 12.7 wt%, N = 6.15 rps, VG = 2, 4.8, 8 and 15 mm/s Solid conc. = 1.8, 5.5, 15 wt%, VG = 10.30 mm/s, N = 6.15 rps

Chapman et al. (1983) T = 0.56 m, H = T , C = T /4 Rewatkar et al. (1991) T = 0.57 m, H = T , C = T /3 Zhu and Wu (2002) T = 0.39 m, H = T , C = T /3

= 2520 kg/m3 ,

2.4 GH and 2 GB memory with each node. Total simulation time for each case was around 120 h. 4. Results and discussions 4.1. Two phase ows It was thought desirable to conrm the validity of the model for the two extreme cases of gasliquid and solidliquid dispersions. For a gasliquid system, particle image velocimetry (PIV) data of Aubin et al. (2004) (see Table 1) have been used for the comparison of radial proles of the mean axial velocity at various axial locations generated by PBTD45 with D/T = 0.5. Further, experimental data of Mishra and Joshi (1991) have been used for the comparison of radial gas hold-up proles at various axial locations. Fig. 1 shows the schematic representation of stirred system with stream lines and particles trajectories. The above simulations have been carried out using the modied Brucato drag model (Khopkar et al., 2006) with an appropriate grid resolution. Fig. 2 (z/T = 0.19 (A), 0.31 (B), 0.49 (C) and 0.65 (D)) shows an excellent agreement between the predictions and experimental data. In case of gas hold-up proles, the present model is quite successful in predicting the local gas hold-up values at various axial locations of z/T = 0.19, 0.31 and 0.49 (Fig. 3). For solidliquid systems, Pinelli et al. (2001) drag model was used. From Fig. 4 it can be seen that the CFD predictions of the axial solid concentration proles are in good agreement with the experimental measurements of Barresi and Baldi (1987). The impeller geometry and reactor details are given in Table 1. The inuence of solid particle concentration on the mean liquid velocity has been studied using CFD and compared with the experimental data of Angst et al. (2004). It can be seen in Figs. 4BD that the present model has predicted the decrease in the mean liquid velocity (at z/T = 0.40) with an increase in the solid concentration (1, 2 and 3 vol%) which is in good agreement with the experimentally measured velocities. 4.2. Three phase ows In Section 4.1, the validity of the CFD model was shown for gasliquid and solidliquid systems. However, for the case of three phase systems, experimental data on gas and solid hold-up proles are not available in the published literature. Therefore, the present simulations have been focused on the bulk ow properties such as the critical impeller speed for solid suspension and qualitative features of liquid circulation. Simulations have been performed over a wide range of experimental conditions as summarized in Table 2.

Fig. 1. Schematic set-up: (A) streamlines; (B) particle trajectories.

7188

B.N. Murthy et al. / Chemical Engineering Science 62 (2007) 7184 7195

0.1 0.05

0.3 0.2 0.1

DIMENSIONAL MEAN AXIAL VELOCITY, W/Utip, (-)

0 0 -0.05 -0.1 0.2 0.4 0.6 0.8 1

0 -0.1 -0.2 -0.3 0 0.2 0.4 0.6 0.8 1

-0.15 0.3 0.2 0.1 0 0 -0.1 0.2 0.4 0.6 0.8

-0.4 0.06 0.04 0.02 0 1 -0.02 -0.04 0 0.2 0.4 0.6 0.8 1

-0.2 -0.3

-0.06 -0.08 NORMALISED RADIAL COORDINATE, r/R (-)

Fig. 2. Comparison between the simulated and experimental proles of the dimensionless mean axial velocity for PBTD45 at various axial levels. (A) z/T = 0.19 m; (B) z/T = 0.31 m; (C) z/T = 0.49 m; (D) z/T = 0.65 m: experimental; CFD predictions.

4.2.1. Gross ow eld The gasliquidsolid ows generated by DT, PBTD45 and PBTU45 impellers have been computed for a supercial gas velocity of 2 mm/s, a solid loading of 3.4 wt% and for the impeller speeds (N ) above the respective critical speeds for gas dispersion (7.5, 6.5 and 8 rps). The predicted liquidvelocity vectors have been depicted in Figs. 5AC, for all the three impellers. It can be observed that the present CFD model is able to capture all the qualitative ow features generated by various impellers. Fig. 5A shows that, in the case of DT impeller, the liquid ow leaving the impeller travels in the radial direction and near the wall splits into two streams. Each stream creates a circulation loop, one below and one above the impeller. Only a part of the energy supplied by the impeller, which is associated with lower loop, is available in the bottom region for performing various functions such as solid suspension. PBTD impeller (Fig. 5B) generates one circulation loop where the ow leaving the impeller is downward toward the bottom of the tank, and is directly available for the suspension. PBTU impeller generates one circulation loop and the liquid ow moves upwards toward the surface of liquid and turns down to the bottom. The length of the liquid path and the number of direction changes are greater in the case of PBTU and DT as compared that for PBTD ow. As a result, the energy associated with the PBTD ow (in the bottom region) is much higher than the DT and PBTU ows and hence

the turbulence intensity for the PBTD impeller is also relatively high. Therefore, a PBTD impeller is relatively more efcient under otherwise identical design and operating parameters (T , D, C, H, P /V , VG , etc.). The CFD predictions of the overall gas hold-up have been compared with the experimental data of Chapman et al. (1983) in Table 3. A satisfactory agreement can be seen between the CFD predictions and the experimental measurements. The simulated gas hold-up distribution in the mid-plane between the two bafes shows gas accumulation in the low-pressure region behind the impeller blades forming the so-called gas cavities. 4.2.2. Solid suspension studies The suspension of solid particles in a three phase gasliquidsolid system has been studied by Zlokarnik and Judat (1969), Queneau et al. (1975), Subbarao and Taneja (1979), Wiedmann and Efferding (1980), Chapman et al. (1983), Wong et al. (1987), Rewatkar et al. (1991), Zhu and Wu (2002), and Dohi et al. (2004). These studies have been critically reviewed by Kasat and Pandit (2005). It was thought desirable to undertake systematic CFD simulation of three phase stirred dispersions. The simulations have been validated by comparing the CFD predictions and the experimental measurements of critical impeller speed for solid suspension over a wide range of design and operating conditions (Table 1). In the

B.N. Murthy et al. / Chemical Engineering Science 62 (2007) 7184 7195

7189

0.4

dened as = 1 n
n 1

0.3

S 1 S

(15)

0.2

0.1

0 0 (-) 0.4
G

0.2

0.4

0.6

0.8

0.3

0.2

0.1

0 0 0.4 0.2 0.4 0.6 0.8 1

0.3

0.2

0.1

0 0 0.2 0.4 0.6 0.8 NORMALISED RADIAL COORDINATE, r/R (-) 1

Fig. 3. Comparison between the simulated and experimental proles of the fractional gas hold-up for PBTD45 at various axial levels. (A) z/T = 0.19 m; (B) z/T = 0.31 m; (C) z/T = 0.49 m: experimental; CFD predictions.

where n is the number of sampling locations used for measuring the solid phase hold-up. The increase in the degree of homogenization (better suspension quality) is manifested as the reduction of the value of standard deviation. On the basis of the quality of the suspension, the range of the standard deviation has been broadly divided into three ranges (Oshinowo and Bakker, 2002). For uniform (homogeneous) suspensions, the value of the standard deviation is found to be smaller than 0.2 ( < 0.2). However, for the just suspension condition, the value of the standard deviation lies between 0.2 and 0.8 (0.2 < < 0.8), and for an incomplete suspension, > 0.8. The CFD simulations were performed to calculate the values of the standard deviation using Eq. (15) for all the three impeller designs (VG = 4 mm/s, 3.4 wt% solid loading, 180 m and pipe sparger). In the present study, the standard deviation was calculated using the values of S stored at all computational cells. Fig. 6 shows the variation of the standard deviation with respect to impeller speed. It can be noted that there is a sharp reduction in the standard deviation as the impeller speed approaches NCS . It can be noted that, at the experimentally measured values of NCS (6.75 rps (PBTD), 10.25 rps (DT) and 9.3 rps (PBTU)), the corresponding predicted values of are 0.76, 0.76 and 0.78, respectively. Therefore, the critical impeller speed for suspension was considered to be achieved when the predicted value of was 0.75. Further, with an increase in the impeller speed, the value of decreases rather slowly. This means, at a constant loading of solid particles and supercial gas velocity, when the impeller speed is gradually increased, beyond NCS , more particles get suspended. The position of solids on the tank bottom (N > NCS ) depends on the impeller design. It can be clearly seen from Fig. 7 that in the case of DT and PBTU impellers, the particles are suspended from an annular space around the center of the tank bottom, whereas for a PBTD impeller suspension occurs from the periphery of tank bottom. It is evident that the impeller speed required for suspension by a PBTD impeller (in fact P /V ) is much lower than required by PBTU and DT impellers. Further, for DT and PBTU impellers the present CFD model predicts a signicant quantity of unsuspended particles present on the tank bottom. This further shows that, at NCS , PBTD45 is more efcient than the DT and PBTU impellers. 4.2.2.1. Effect of impeller design. Earlier it has been shown that DT, PBTD and PBTU impellers generate different ow patterns, and hence, offer different efciencies for the suspension operation. In order to understand the quantitative role of the impeller design, CFD simulations have been carried out for the three impeller designs and at different impeller speeds. For brevity, qualitative results are shown in Figs. 7AC at the critical impeller speeds. Figs. 7DF show the axial concentration proles with respect to impeller rotational speed for all the three impeller designs, where the axial solid concentrations

EulerianEulerian approach as used in this work, it is difcult to incorporate Zwieterings criterion in the CFD simulation of critical impeller speed for solid suspension. Therefore, we have extended the method proposed by Bohnet and Niesmak (1980) for solidliquid system, which has been based on the value of standard deviation of solid concentration. The same methodology has successfully been employed by Oshinowo and Bakker (2002) and Khopkar et al. (2006). Bohnet and Niesmak (1980) quantied the suspension quality using the standard deviation

FRACTIONAL GAS HOLD-UP,

7190

B.N. Murthy et al. / Chemical Engineering Science 62 (2007) 7184 7195

0.3 DIMENSIONLESS MEAN AXIAL VELOCITY, W/Utip (-) 1.0 0.8 0.6 z/T 0.4 0.2 0.0 0.0 0.5 C/Cavg 1.0 1.5 0.2 0.1 0 0 -0.1 -0.2 -0.3 NORMALISED RADIAL COORDINATE, r/R (-) 0.2 0.4 0.6 0.8 1

0.4 DIMENSIONLESS MEAN AXIAL VELOCITY, W/Utip (-) 0.3 0.2 0.1 0 -0.1 -0.2 -0.3 NORMALISED RADIAL COORDINATE, r/R (-) 0 0.2 0.4 0.6 0.8 1 DIMENSIONLESS MEAN AXIAL VELOCITY, W/Utip (-)

0.4 0.3 0.2 0.1 0 -0.1 -0.2 -0.3 NORMALISED RADIAL COORDINATE, r/R (-) 0 0.2 0.4 0.6 0.8 1

Fig. 4. (A) Comparison of experimental and predicted axial solid concentration proles at Nj s of different particle sizes ( : dp = 100.177 m; : dp = 208.250 m; : dp = 417.500 m) for 4-PBTD. (B) Comparison of experimental and predicted dimensional mean axial proles with mean dispersed phase volume fractions of (B) 1 vol%, (C) 2 vol%, (D) 3 vol%, stirrer speed 1000 s1 : experimental, CFD predictions.

Fig. 5. Axial velocity (mm/s) vectors for liquid in presence of gas and solid in the mid-plane between two bafes: (A) DT; (B) PBTD45; (C) PBTU45.

B.N. Murthy et al. / Chemical Engineering Science 62 (2007) 7184 7195


Table 3 Experimental and predicted values of overall gas hold-up
S. no. Impeller speed (s1 ) 3.3 4 5 Gas sparging rate (m3 /s) 0.57 1.14 2.32 % Gas hold-up (G ) Experimental 1 2 3 1.7 2.8 5.6 Predicted 1.5 2.5 5

7191

in the particle size. With increasing particle size the settling velocity increases and there is a decrease in the homogeneity of the suspension. Therefore, higher average liquid velocity is required to suspend the particles. 4.2.2.3. Effect of solid loading. CFD simulations were performed to investigate the effect of solid loading on the critical impeller speed for solid suspension. For this, particles of diameter 180 m particles with solid loadings of 3.4, 6.6 and 12.7 wt%, respectively, have been considered (with a PBTD45 impeller, pipe sparger, VG = 8 mm/s and at various impeller rotational speeds). Fig. 9 shows a fairly good agreement between the CFD predicted and the experimentally measured values of NCS . The present model is able to simulate (results are not given for brevity but for two phase ows see Figs. 4BD) the decrease in liquid ow with an increase in the solid loading. This is because some of the impeller energy dissipates at the solidliquid interface. 4.2.2.4. Effect of supercial gas velocity. In order to study the capability of the present CFD model to simulate the effect of supercial gas velocity on the distribution of solids, CFD simulations have been performed over a wide range of supercial gas velocity. Fig. 10 shows a good agreement between the CFD predictions and the experimental data on NCS for all the impeller designs and at VG = 2, 4 and 8 mm/s (pipe sparger 60 mm). The qualitative distribution of solids in the absence of gas at the critical impeller speed of 6.5 rps (Rewatkar et al., 1991) is shown by the contours of S in Fig. 11A. On the introduction of gas (VG = 2 mm/s) experimentally it has been reported that the solid particles settle down due to a decrease in the circulation velocity as well as the liquid turbulence. As a result there is a reduction in the solids cloud height. The same behavior has been qualitatively well predicted by the present CFD model as shown in Fig. 11B. Further, the predicted axial dimensionless concentration proles with respect to increasing impeller speed are shown in Fig. 11C. It shows that with increasing impeller speed the amount of particles present at the bottom of the reactors has decreased which has been in agreement with experimentally reported observations. However, the increased impeller speed has marginal inuence on the solid distribution in the top zone of the reactor. In order to model the suspension capabilities of all the three impellers at NCS , the simulations of solid suspension were performed using the present model at supercial gas velocities of 2,4,6,8 and 10 mm/s. The predicted values of standard deviation are shown in Fig. 12. Experimentally it was observed that the effect of VG on the amount of solids suspended depends upon the type of impeller because the reduction in power with increasing VG varied with the type of impeller. It can be seen that the present CFD model was able to predict a gradual increase in with an increase in VG for PBTD impeller and hence relatively less particles tend to settle down at the bottom of the reactor. Whereas with DT and PBTU impellers there is a remarkable rise in on the introduction of gas and more and more solid particles start getting settled with increasing VG . This essentially means that PBTD impeller is

2.4 2 1 1.6 3 1.2 0.8 0.4 0 4 5 6 7 8 9 10 11 12 13 IMPELLER ROTATIONAL SPEED, N (1/s)


Fig. 6. CFD predicted values of the standard deviation with respect to impeller rotational speed: (1) PBTD; (2) DT; (3) PBTU (VG = 4 mm/s, 3.4 wt% solid loading, 180 m and pipe sparger).

STANDARD DEVIATION, j (-)

1 PBTD (CFD) 2 DT (CFD) 3 PBTU (CFD)

are made dimensionless by dividing local solid concentration by the average solid concentration. It can be noted that with an increase in the impeller rotational speed the amount of solid particles present at the bottom of the reactors has decreased. However, the increased impeller speed has not much inuence 1 on the solids distribution in top 4 th of the reactor. The values of the standard deviation have been calculated using Eq. (17). Predictions of critical impeller speeds using CFD (when = 0.75) for all the impeller designs are compared with the experimentally measured critical impeller speed. Table 4 shows an excellent agreement between the values of NCS for PBTD, DT and PBTU for given reactor geometry (pipe sparger), solid loading (3.4 wt%) and particle diameter (180 m) at 6.5, 9, and 11 rps. 4.2.2.2. Effect of particle size. The critical impeller speed for solid suspension also depends upon the particle size. Therefore, it was thought desirable to study the predictive capabilities of the present CFD model for various particle sizes. The simulations have been carried out for the three particle diameters, i.e., 180, 340 and 700 m with PBTD45, DT and PBTU45 impellers, pipe sparger, 6.6 wt%, VG = 4.8 mm/s and at various impeller speeds. Fig. 8 shows a good agreement between the CFD predictions and the experimentally measured data for all the impeller designs. Further, it conrms the fact that, for a given tank and impeller conguration and for xed set of operating conditions, uniformity of solids increases with a decrease

7192

B.N. Murthy et al. / Chemical Engineering Science 62 (2007) 7184 7195

0.025

0.018

0.0125

0.06

0.00 1 0.8 500 rpm 550 rpm 600 rpm 0.6 0.4 0.2 0 0 1 2 3 0 0.4 0.8 1.2 1.6 2 360 rpm 420 rpm 480rpm

1 0.8 0.6 0.4 0.2 0

1 0.8 0.6 0.4 0.2 0 0 0.4 0.8 1.2 1.6 2 620 rpm 680 rpm 580 rpm

Fig. 7. Effect of impeller type on solid concentration distribution for (A) DT; (B) PBTD; (C) PBTU at NCS (Rewatkar et al., 1991) by CFD simulations (dp = 180 m, p = 2520 kg/m3 , 6.6 wt%, VG = 2 mm/s (pipe sparger)).

CRITICAL IMPELLER SPEED FOR SOLID SUSPENSION, NCS (1/s)

Table 4 Effect of impeller type on solid concentration distribution for (dp = 180 m, 3 p = 2520 kg/m , 6.6 wt%, VG = 2 mm /s (pipe sparger))
S. no. Impeller type Critical impeller speed for off-bottom suspension (NCS ) (1/s) Experimental 1 2 3 PBTD DT PBTU 6.5 9.2 11.1 Simulated 6.5 9.5 11.3

16 14 12 10 8 6 4 2 0 0 200 400 600 AVERAGE PARTICLE SIZE, dp (micron) 800 CFD PBTD (Exp) DT (Exp) PBTU (Exp)

still efcient than DT and PBTU on the introduction VG at a given impeller speed. 4.2.2.5. Design of sparger. The ow generated by all the impeller types get modied by the presence of gas and also the way in which the gas is sparged into tank, i.e., the sparger design. Therefore, in order to model the effects of sparger design on NCS the present CFD model is used. The simulations of solid suspension were performed for the pipe sparger and ring spargers of different diameter (with a PBTD45,

Fig. 8. Comparison of experimental and predicted critical impeller speeds for different impeller designs. PBTD; DT; PBTU (experimental), CFD predictions (dp = 180 m, p = 2520 kg/m3 , 6.6 wt%, VG = 4.8 mm/s (pipe sparger)).

VG = 8 mm/s, dp = 180 m, sparger is located at 100 mm from the bottom and solid loading of 3.4 wt%). It is evident from Table 5 that the predicted NCS values agree fairly well with the

B.N. Murthy et al. / Chemical Engineering Science 62 (2007) 7184 7195

7193

CRITICAL IMPELLER SPEED FOR SOLID SUSPENSION, NCS (1/s)

10

experimental values for all the sparger designs. It can be noted that the ring sparger provides lower value of NCS compared to that for the pipe sparger. Further, sparger having large ring diameter is found to give the lowest value of NCS compared to
2.8 STANDARD DEVIATION, (-)

6 1 PBTD (CFD) 2 DT (CFD) 3 PBTU (CFD) 1 1.6 1.2 0.8 0.4 0 2

4 CFD PBTD (Exp)

2.4 2

0 0 5 10 15 SOLID LOADING, X (%, wt.)

Fig. 9. Effect of solid loading on the critical impeller speed for PBTD (dp =180 m, p =2520 kg/m3 , VG =8 mm/s (pipe sparger)): experimental, CFD predictions.

CRITICAL IMPELLER SPEED FOR SOLIDSU SPENSION, NCS (1/s)

11

2 4 6 8 10 SUPERFICIAL GAS VELOCITY,VG (mm/s)

12

10

Fig. 12. CFD predicted values of the standard deviation at NCS with respect to supercial gas velocity: (1) PBTD; (2) DT; (3) PBTU (3.4 wt%, 180 m and pipe sparger).

8 CFD PBTD (Exp) DT (Exp) PBTU (Exp) 0 2 4 6 8 SUPERFICIAL GAS VELOCITY, VG (mm/s) 10

Table 5 Design of sparger on solid concentration distribution for PBTD (dp = 180 m, 3 p = 2520 kg/m , 6.6 wt%, VG = 8 mm /s)
S. no. Sparger type Critical impeller speed for off-bottom suspension (NCS ) (1/s) Experimental 1 2 3 4 5 SP60 SR95 SR190 SR152 SR420 7.8 7.5 7.25 7 6.5 Simulated 8 7.75 7.5 7.2 6.5

Fig. 10. Comparison of experimental and predicted critical impeller speeds for DT; PBTU (experimental), various supercial gas velocities. PBTD; CFD predictions (dp = 180 m, p = 2520 kg/m3 , 6.6 wt%, VG = 2, 4 and 8 mm/s (pipe sparger)).

5.20e-02

5.20e-02

1 0.8 390rpm 550rpm 0.6 z/T 450rpm

4.00e-02

4.00e-02

2.95e-02

2.95e-02

0.4 0.2 0 0 0.4 0.8 1.2 C/Cavg 1.6 2

1.85e-02

1.85e-02

0.00e-00

0.00e-00

Fig. 11. Effect of supercial gas velocity on solid concentration distribution for PBTD by CFD simulations at 6.5 rps (Rewatkar et al., 1991) (dp = 180 m, 3 p = 2520 kg/m , 6.6 wt%, (ring sparger, 0.095 mm o.d.)). (A) VG = 0 mm /s; (B) VG = 2 mm /s; (C) dimensionless axial concentration proles at 390, 450 and 550 rpm.

7194

B.N. Murthy et al. / Chemical Engineering Science 62 (2007) 7184 7195

Fig. 13. Effect of sparger on solid concentration distribution for PBTD at NCS (Rewatkar et al., 1991) by CFD simulations (dp = 180 m, p = 2520 kg/m3 , 6.6 wt%, VG = 10 mm/s): (A) pipe sparger (0.06 m); (B) ring sparger (0.095 m o.d.); (C) ring sparger (0.42 m o.d.).

the smaller one which are in good agreement with the experimental ndings. Experimentally it has been observed that the ring sparger with diameter 2D (SR420) sparges the gas along the periphery and generates the ow pattern in the same direction as that generated by the impeller action of PBTD. Therefore, it helps to maintain better suspension of the solid particles throughout the reactor. Figs. 13AC show the contour plots of S at NCS in the mid-bafe plane for the three sparger designs (for brevity, SP60, SR95 and SR420). The concentration distribution is more uniform with SR420 compared to SP60 and SR95. 5. Conclusions (1) In the present work, three phase stirred suspension has been simulated using FLUENT 6.2 CFD software. The Eulerian multi-uid model along with the standard k turbulence model has been used to simulate gasliquid, solidliquid and gasliquidsolid dispersions. (2) A very good agreement was found between the predicted and the experimental velocity and G proles in gasliquid dispersions. (3) A very good agreement was also found between the predicted and experimental proles of S over a wide range of impeller speed and the impeller design. (4) By using the concept proposed by Oshinowo and Bakker (2002) a value of standard deviation ( = 0.75) has been suggested for the prediction of critical impeller speed for solid suspension. The suggested value of holds for different impeller designs and over a wide range of particle size, solid loading and supercial gas velocity. (5) For three phase dispersions, the predicted critical impeller speeds have been compared with the experimental results of Chapman et al. (1983), Rewatkar et al. (1991) and Zhu and Wu (2002) over a wide range of solid loading

(0.3415 wt%), for different impeller designs (Rushton turbine (RT), pitched blade down and upow turbines (PBT45)), solid particle sizes (1801000 m) and for various supercial gas velocities (010 mm/s). A very good agreement was observed in all these cases.

Notation C , C 1, C C Cavg CD CDO dB dp D g H k MD N NCS Nj s p P PK t T u V VG w z


2

turbulence model constants solid concentration, kg/m3 average solid concentration, kg/m3 drag coefcient in turbulent liquid drag coefcient in still liquid bubble diameter, m particle diameter, m impeller diameter, m acceleration due to gravity, 9.8 m/s2 liquid height, m turbulent kinetic energy, m2 /s2 drag force per unit area, N/m3 impeller rotation speed, s1 critical impeller speed for solid suspension in gasliquidsolid system, rps critical impeller speed for just suspension, rps pressure, Pa power consumption, W turbulence production, kg/m1 s3 time, s tank diameter, m average velocity, m/s volume of the reactor, m3 supercial gas velocity, m/s width of the impeller blade, m axial co-ordinate direction, m

B.N. Murthy et al. / Chemical Engineering Science 62 (2007) 7184 7195

7195

Greek letters Kolomogoroff eddy size, m turbulent energy dissipated per unit mass, m2 /s3 volume fraction viscosity, kg/m s density of uid, kg/m3 turbulent Prandtl number for the dissipation rate turbulent Prandtl number for the turbulent kinetic energy

Subscripts eff G L LG LS S r t tip effective gas phase liquid phase liquidgas liquidsolid solid phase radial co-ordinate direction turbulent at the tip of the impeller blade

Acknowledgment Mr. B.N. Murthy and Dr. R.S. Ghadge gratefully acknowledge the nancial support during this work by Department of Atomic Energy (DAE), Government of India. References
Angst, R., Kraume, M., Ritter, J., 2004. Particle distribution and velocity in stirred vessels: experimental results and CFD-simulations. In: Proceedings of the Third International Symposium on Two-Phase Flow Modelling and Experimentation, Pisa, pp. 221229. Aubin, J., Sauze, N.L., Bertrand, J., Fletcher, D.F., Xuereb, C., 2004. PIV measurements of ow in an aerated tank stirred by a down- and an uppumping axial ow impeller. Experimental Thermal and Fluid Science 28, 447456. Barresi, A., Baldi, G., 1987. Solid dispersion in an agitated vessel. Chemical Engineering Science 42, 29492956. Bohnet, M., Niesmak, G., 1980. Distribution of solids in stirred suspension. German Chemical Engineering 3, 5765. Chapman, C.M., Nienow, A.W., Cooke, M., Middleton, J.C., 1983. Particlegasliquid mixing in stirred vessels, part III: three phase mixing. Chemical Engineering Research and Design 60, 167181. Dohi, N., Takahashi, T., Minekawa, K., Kawase, Y., 2004. Power consumption and solid suspension performance of large-scale impellers in gasliquidsolid three-phase stirred tank reactors. Chemical Engineering Journal 97, 103114. Elgobashi, S.E., Abou-arab, T.W., 1983. A two-equation turbulence model for two-phase ows. Physics of Fluids 26, 931938.

FLUENT 6.2., 2005. Users Manual to FLUENT 6.2. Fluent Inc. Centrera Resource Park, 10 Cavendish Court, Lebanon, USA. Hinze, J.H., 1975. Turbulence. McGraw-Hill, New York. Kasat, G.R., Pandit, A.B., 2005. Review on mixing characteristics in solidliquid and solidliquidgas reactor vessels. The Canadian Journal of Chemical Engineering 83, 618642. Khopkar, A.R., Rammohan, A.R., Ranade, V.V., Dudukovic, M.P., 2005. Gasliquid ow generated by a Rushton turbine in stirred vessel: CARPT/CT measurement and CFD simulations. Chemical Engineering Science 60, 22152229. Khopkar, A.R., Kasat, G.R., Pandit, A.B., Ranade, V.V., 2006. CFD simulation of mixing in tall gasliquid stirred vessel: role of local ow patterns. Chemical Engineering Science 61, 29212929. Ljungqvist, M., Rasmuson, A., 2001. Numerical simulation of the two phase ow in an axially stirred reactor. Transactions of the Institution of Chemical Engineers 79, 533. Mishra, V.P., Joshi, J.B., 1991. LDA measurements of ow in stirred gasliquid reactors. In: Proceedings of the Seventh European Congress on Mixing, Brugge-Belgium, pp. 217224. Montante, G., Micale, G., Magelli, F., Brucato, A., 2001. Experiments and CFD predictions of solid particle distribution in a vessel agitated with four pitched blade turbines. Transactions of the Institution of Chemical Engineers 71, 10051010. Nigam, K.D.P., Schumpe, A., 1996. Three-Phase Sparged Reactors. Gordon and Breach Science Publishers. Oshinowo, I.M., Bakker, A., 2002. CFD modelling of solid suspensions in stirred tanks. Symposium on Computational Modelling of Metals, Minerals and Materials, TMS Annual Meeting, Seattle, WA, pp. 234242. Pinelli, D., Nocentini, M., Magelli, F., 2001. Solids distribution in stirred slurry reactors: inuence of some mixer congurations and limits to the applicability of a simple model for predictions. Chemical Engineering Communications 188, 91107. Queneau, P.B., Jan, R.J., Richard, R.S., Lowe, D.F., 1975. Turbine miner fundamentals and scale-up at port nickel. Metallurgical Transactions B 6B, 149155. Rewatkar, V.B., Raghava Rao, K.S.M.S., Joshi, J.B., 1991. Critical impeller speed for solid suspension in mechanical agitated three-phase reactors. 1. Experimental part. Industrial and Engineering Chemistry Research 30, 17701784. Subbarao, D., Taneja, V.K., 1979. Three phase suspension in agitated vessels. In: Proceedings of the Third European Conference on Mixing, BHRS, Carneld, UK, pp. 229240. Viollet, P.L., Simonin, O., 1994. Modelling dispersed two-phase ows: closure, validation and software development. Applied Mechanical Reviews 47 (6), S80S84. Wiedmann, J., Efferding, L.E., 1980. Experimental investigation of suspension, dispersion, power, gas hold-up and ooding characteristics in stirred gasliquidsolid systems. Chemical Engineering Communications 6, 245256. Wong, C.W., Wang, J.P., Haung, S.T., 1987. Investigations of uid dynamics in mechanically stirred aerated slurry reactors. The Canadian Journal of Chemical Engineering 65, 412419. Zhu, Y., Wu, J., 2002. Critical impeller speed for suspending solids in aerated agitation tanks. The Canadian Journal of Chemical Engineering 80, 16. Zlokarnik, N.W., Judat, P., 1969. Tubular and propeller stirrersan effective stirrer combination for simultaneous gassing and suspending. Chemie Ingenieur Technik 41 (23), 12701277.

You might also like