Sie sind auf Seite 1von 17

www.catchword.com=titles=02638762.

htm

02638762/02/$23.50+0.00 # Institution of Chemical Engineers Trans IChemE, Vol 80, Part A, July 2002

CFD STUDY OF POWER AND MIXING TIME FOR PADDLE MIXING IN UNBAFFLED VESSELS
S. MURTHY SHEKHAR and S. JAYANTI
Department of Chemical Engineering, IIT-Madras, Chennai, India

FD-based computations of the ow eld, power consumption and mixing time are presented for a mechanically stirred eight-blade paddle impeller in an unbaf ed vessel over a range of Reynolds numbers covering laminar, transitional and turbulent ow regimes. The ow eld calculations were performed using the sliding mesh technique to account for the motion of the impeller, and mixing time studies were done using a simulated tracer injection experiment. The effect of grid density and the choice of the turbulence model were investigated. The results are compared with ow eld data from Dong et al.1, and power and mixing time correlations from the literature and show satisfactory agreement. It is shown that the product of mixing time and rotational speed remains constant for paddle impellers for laminar ow and that the use of a low Reynolds number turbulence model is necessary for good prediction of mixing time in the transitional ow. Keywords: mixing; power; mixing time; CFD simulations; sliding mesh technique; turbulence modelling; paddle impeller.

INTRODUCTION Mixing is an important unit operation, accomplished either by rotating impellers or by jets of liquid. The essential requirement of the mixers is to bring together two or more uids which are initially separated. The optimum design of a stirred tank depends on the desired production rate, properties of the uid to be agitated, choice of tank and impeller geometry, rotational speed and location of uid addition and removal. This requires a detailed knowledge of the mixing characteristics of stirred tanks. A number of systematic studies of mixing have been conducted over the past several decades (see, for example, references 2,3). The emphasis of early studies was, principally, on three aspects: impeller geometry, power consumption and mixing time for a given mixing duty. A large number of power and mixing time measurements and correlations are available in the literature for impellers of various geometries and for various uids. These correlations have invariably been obtained based on laboratory scale measurements and their scaling up to industrial scale mixing devices has always been a matter of concern. In recent years, a more fundamental approach has been taken to the understanding of the mixing process through detailed studies of the ow pattern created by an impeller. Computational Fluid Dynamics (CFD) techniques are being increasingly used as a substitute for experiments to obtain the detailed ow eld for a given set of uid, impeller and tank geometries1,411. One advantage with CFD-based prediction methods is that these do not have scaling up or scaling down problems as these solve the fundamental equations governing uid ow. Still, some approximation of the physical phenomena is often required even in CFD 482

simulations. Examples of these are the phenomenological models for turbulence12; rheological models for nonNewtonian uids13 and impeller-speci c boundary conditions being used in simulating mixing in some of the earlier studies5,6,14. However, more recent studies have addressed all three aspects. Jenne and Reuss15 have recently conducted an assessment of the k-e turbulence for baf ed stirred tank reactors and proposed modi cation of the ChenKim model16. Ciofalo et al.8 and Armenante et al.17 have used second order closure turbulence models and found better agreement with measured data. Simulations of high viscosity power-law uids using anchor and helical ribbon impellers have been reported in the literature18,19. While empirical relations and measured velocity and other parameters are still being used for the speci cation of the boundary conditions on the impeller5,6,14,15, there is now a tendency to do away with this empiricism and resolve the ow eld over the impeller as part of the calculation4,8,10,11,18. Good to satisfactory agreement with measured ow eld has been reported in several studies. There is, thus, a vast amount of CFD literature on the subject of mixing in stirred tank reactors. However, most of these studies have focussed on the prediction of the velocity eld alone. Some studies7,8,17 have reported a few power computations. While Ranade et al.20 have studied mixing for a disc turbine, no systematic computational study of mixing time appears to have been attempted yet for low viscosity uids. Also, many of the reported calculations of paddle=pitched blade impellers are in the high Reynolds number NRe > 10;000 regime and none have been reported in the laminar regime NRe < 10: While paddletype impellers may not nd much use under laminar ow

CFD STUDY OF POWER AND MIXING TIME conditions, studies in this range would be of academic interest as the errors induced due to turbulence modelling would be eliminated therein. Finally, the simulation of the transition regime 10 < NRe < 1000 requires special attention as the ow is neither fully laminar nor fully turbulent. It would be of interest to see how CFD simulations perform in this regime. In the present study, these three issues are addressed. Mixing of a Newtonian uid by a paddle-type impeller in an unbaf ed vessel is taken as a case study and calculations of the velocity eld have been performed with the sliding mesh method9,11 for the experimental case investigated by Dong et al.1. These have then been extended to the calculation of the power consumption and mixing time over a range of Reynolds numbers not hitherto explored. The details of these calculations and the results are discussed below. MATHEMATICAL FORMULATION Governing Equations The basic equations solved in a mixing calculation are those describing the ow of uids, namely, conservation of mass, momentum and energy. Since mixing usually involves liquids, the uid can be taken to be incompressible. For cases where there is no heating or cooling involved (neglecting for the time being the temperature rise due to viscous dissipation), the ow is governed by the simpli ed set of continuity and Navier-Stokes equations for an incompressible Newtonain uid: continuity: @ui 0 @xi @ui @ui uj 1 @p @2 u n 2i r @xi @t @xj @xj 1 2

483

mediate regime have been performed using the low Reynolds number k-e model23 as discussed below. The derivation of the standard k-e turbulence is discussed in detail in many books21 and is not reproduced here. It is suf cient to note that for turbulent ows, the variables appearing in equations (1) and (2) above would be timeaveraged and would be supplemented with two additional transport equations, one each for the turbulent kinetic energy (k) and its dissipation rate (e). These equations are given, for example, in Reference 20. The standard k-e model is useful only for fully turbulent ows, for example, for ow in a pipe at a Reynolds number of 50,000. For ows induced by an impeller, it is likely that the region near the impeller is turbulent for suf ciently high impeller speeds. However, at distances well away from the impeller, large velocity gradients, which are the main source of turbulent energy, may not be present, and turbulent kinetic energy may be dissipated resulting in a relaminarization of ow. Hence, there is a possibility that the ow near the impeller is turbulent while it may be laminar well away from it. The use of standard turbulence models based on the assumption of high Reynolds number and the consequent absence of the effect of laminar viscosity may be wrong in such a case. Jayanti and Hewitt24 reported success in predicting ows in which part of the ow domain is laminar using the low Reynolds number version of the standard k-e model23. This model was originally developed to take account of relaminarization of turbulent ow in regions of high acceleration and was used by Jayanti and Hewitt24 for the calculation of a wavy, thin lm ow under large shear. In the present study, additional calculations have been performed with this model to account for possible laminarization of ow well away from the impeller. Two major modi cations are present in the low Reynolds k-e model as compared to the standard k-e model: (i) the effect of molecular viscosity is included in the model, and the turbulent kinetic energy (k), and its rate of its dissipation, are now made to become less as the wall is approached, with their being totally suppressed at y < 5; (ii) the conservation equation for e is slightly modi ed to enable the resolution of the ow right through to the wall. A much closer grid spacing near the wall is required for this simulation than the standard k-e model. Details of this model and the rationale behind the above changes are explained in detail by Jones and Launder23.

momentum:

Here ui is the velocity in the ith direction, r is the density, p is the pressure and n is the kinematic viscosity of the uid and the summation convention for repeated indices is assumed. In the above equations, the time-dependent terms are retained as mixing time can be estimated directly from time-dependent simulations, as will be shown later. For turbulent ow, the above set of equations is still valid but will have to be solved with a very ne spatial and temporal resolution (the so-called direct numerical simulation (DNS) of turbulent ows) to obtain the true variation of the velocity eld. This is often beyond the scope of many studies and often phenomenological models of turbulence are used. Here, the governing equations are time-averaged21 and the resulting set of time-averaged equations are supplemented with additional transport equations to model the Reynolds stresses, i.e., the cross-correlations between uctuating velocity components. Many such turbulence models exist, with the standard k-e model being the most often used for industrial calculations. For impeller-driven ows, ows having an impeller Reynolds number NRe rNd 2 =m greater than 10 are considered to be in the transitional or turbulent regime22. Therefore, in the present study the standard k-e model is used for NRe > 500; while laminar ow is imposed for NRe < 10: Calculations for the interTrans IChemE, Vol 80, Part A, July 2002

Representation of the Impeller A number of strategies can be used to deal with the movement of the impeller blades. These include: (i) use of a momentum source term7; (ii) solving the governing equations in a rotating coordinate system2,25; (iii) the sliding mesh approach9,11. The rst of these requires some empirical treatment of the momentum exchange between the impeller blades and the bulk uid and, at best, requires calibration. The second approach has the advantage of permitting a steady state calculation even though the impeller is moving but cannot deal with an unsteady state calculation, as required, for example, for tracer study to simulate mixing26. It also requires that there exists a circular

484

MURTHY SHEKHAR and JAYANTI Estimation of Power Required The fact that the ow eld around the impeller is resolved enables the estimation of the power directly from a calculation of the total torque required to rotate the impeller. Since the ow domain, as well as the enveloping surface, is subdivided into small volumes and areas as part of the CFD solution, the shear stress and the pressure distribution on the impeller blade become available at the end of the calculation, i.e., once a periodic state is reached. The torque on each blade can be readily calculated as: X T Dpi Ai ri 3
i

interface where the ow pattern does not have any angular dependence. Thus, it is strictly to be used for cases where the impeller-baf e interaction is, at most, weak. (This restriction is not a problem in the present study where there are no baf es.) The third method, i.e., the sliding mesh approach, is the most versatile of all but, since it requires a time-dependent calculation, is also the most demanding from a computational resources point of view. Since the present study requires time-accurate solutions for mixing time characterization, the sliding mesh approach is used. The details of the sliding mesh approach have already been discussed in the literature27 and are not repeated here. The basic idea is to employ two grids one of which moves with the impeller while the other is xed to the tank. The meshes interact along a surface of slip and the moving grid is allowed to slide relative to the stationary one. The grid lines are not required to align themselves on the slip surface; a conservative interpolation is used to obtain the ow variables and face uxes across this surface. The principal feature of the sliding mesh method is that the ow eld over the impeller is resolved as part of the solution, and hence, no speci cation other than the geometrical details is necessary to represent the impeller. The required boundary conditions take the form of the no-slip boundary condition on the impeller surface, shaft and other solid boundaries. The rotational speed of the impeller must obviously be speci ed. Other than this, no empirical or impeller-speci c velocity or momentum source conditions need to be speci ed. Thus, there is no empiricism in the model formulation other than what is associated with turbulence modelling and the no-slip condition. The top surface of the liquid is taken as a freeshear surface; however its distortion is not considered and this may introduce some error in the formulation, especially at high Reynolds numbers.

where the summation is carried over all the cells having the blade as one of the boundaries. Here, Dp is the pressure difference between the front (upstream) and the back (downstream) side of the blade at the surface element i, Ai is the projected surface area of the surface element i and ri is the radial distance from the axis of the shaft on which the impeller is mounted. The power required for rotation of the impeller at a steady rotational speed of N revolutions per second for an impeller having m blades is then given by: P 2pNmT

and the power number is computed as NP P=rN 3 d 5 where d is the outer diameter of the impeller. The above procedure of power estimation reduces the computational time requirement, by eliminating the solution of the energy equation to deduce the power from viscous energy dissipation. It is not necessary to calculate the viscous energy dissipation term from the calculated velocity eld. Estimation of Mixing Time In experiments, mixing time is often estimated using a tracer technique2. Typically, the tracer concentration at a point within the tank varies with time, and the time taken for the variation to reduce below a certain level, say, within 5% of the fully-mixed concentration, is taken as the mixing time. The same method can be used in a numerical calculation by injecting a neutrally-buoyant, virtual tracer at a given location. The concentration, or to be exact, its mass fraction, is governed by the following scalar transport equation: @rf H:ruf Gf Hf Sf @t 5

Solution of the Governing Equations In a CFD solution, the ow domain is broken up into a number of contiguous and non-overlapping cells enveloping the whole domain and the ow variables are sought at (the centres of) each of these cells. The governing equations, are therefore, discretized and linearized (in the present case, by a nite volume scheme) resulting in a set of coupled, linear algebraic equations which are then solved using iterative schemes. The various steps in this scheme of solution have become suf ciently standardized that commercially available CFD codes can be used to solve the governing equations. This approach is resorted to in the present case and all the calculations reported here have been done using the CFD code CFX developed by AEA Technology, UK. The version used in the present calculations is a nitevolume based code using structured meshes with multiblock capability. A non-staggered grid in cylindrical coordinates is used to discretize the ow domain. A pressure-correction method of the type Semi-Implicit Method for PressureLinked Equations-Consistent (SIMPLEC28) is used to effect pressure-velocity coupling. All the calculations have been done in a transient mode using an explicit method for time advancement. Since the boundary conditions are held constant with respect to time, the ow eventually reaches a steady state.

where r is the density of the carrier uid, u is its velocity, f is the mass fraction of the tracer, Gf is the tracer diffusion coef cient in the carrier uid, de ned as Gf rDf where Df is the tracer diffusivity. The turbulent diffusion coef cient, Gf of the tracer is due to turbulent eddies in the uid and is calculated using the eddy viscosity hypothesis20 and is derived from the turbulence model. Sf is the source or sink term, if any, of the tracer. As there is no chemical reaction, Sf is zero in the present case. Equation (5) above is solved as an additional scalar transport equation in CFX with Neumann boundary conditions and the mixing time is deduced from the temporal variation of the concentration eld, as explained later. Trans IChemE, Vol 80, Part A, July 2002

CFD STUDY OF POWER AND MIXING TIME Details of the Simulations The system investigated in the present study corresponds to the case investigated by Dong et al.1 experimentally and numerically4 for which data of the ow eld are available but not of the power and the mixing time. The system (see Figure 1(a)) consists of a cylindrical unbaf ed vessel of diameter 0.1 m with an eight-blade at paddle impeller of a diameter of 0.024 m and a width of 0.01 m. The impeller shaft is concentric with the axis of the vessel. The problem has been formulated in cylindrical co-ordinates with three dimensions, the axial, radial and tangential components being u, and w, respectively. Due to the symmetry of the impeller and vessel, calculations were performed for a sector of 45 representing 1=8th of the vessel (see Figure 1(b)) and periodic boundary conditions were imposed in the circumferential directions at the two ends of the sectors. The blade is located at the centre of the sector and typically, the grid has 39 23 12 (henceforth referred to as the low density grid case) nodes in the axial, radial and tangential directions giving the same spatial resolution as used by Dong et al.4. The calculations were repeated with a grid of 78 46 24 (henceforth referred to as the high density grid case), i.e, twice the number of nodes in each direction, to see if this improves the solution. (One set of calculations have been

485

performed with twice the number of cells in the circumferential direction as compared to the high density grid case and the results for the velocity pro les and the power number showed nearly identical results with these two grids containing 86,000 and 172,000 cells.) For simplicity, the impeller is modelled as being as a thin surface with no thickness. The uid is taken to be incompressible with a density of 1000 kg m3 and a kinematic viscosity of 10 3, 105 and 106 m2 s1 for laminar ow NRe < 10; transition ow 10 > NRe < 480 and turbulent ow NRe > 480 conditions, respectively. When a turbulence model is used, the effective viscosity may be higher than this molecular viscosity and will be determined by the local values of k and e. All the calculations were carried out using the commercially available CFX computer code developed by AEA Technology, UK. CFX is a general purpose computer program using a nite volume method based discretization on a non-staggered grid using the Rhie-Chow algorithm29 to avoid chequerboard oscillations associated with the use of non-staggered grids. A suite of discretization schemes, turbulence models etc is available for the user to choose from for a particular problem. In the present study, a secondorder accurate central differencing was used for the discretization of the diffusion terms, while convective ux terms

Figure 1. (a) Schematic diagram of the mixing vessel and the impeller; (b) Details of the simulated geometry indicating the stationary and the sliding parts.

Trans IChemE, Vol 80, Part A, July 2002

486

MURTHY SHEKHAR and JAYANTI

were evaluated using the hybrid scheme which reduces to the rst-order accurate upwind scheme if the mesh Peclet number is greater than 2, and to central differencing otherwise. In the present problem, for all laminar ow cases, the maximum value of Peclet number is of the order of 0.2. For turbulent ows, the Peclet number based on molecular viscosity may be of the order of 10 to 50 near the impeller. However, due to turbulence, the effective viscosity is much higher and the mesh Peclet number is signi cantly less than one. In regions well away from the impeller, although the turbulent viscosity decreases, the velocities are also decreased by an order of magnitude or more. Hence, the discretization is expected be formally second-order accurate. The calculations were started with zero initial velocity condition and were advanced in time with a time step corresponding to a 10 rotation of the impeller. The local and average values of the velocities and stresses were monitored to determine whether or not a steady state was reached. Typically, about 5000 time steps, corresponding to about 150 revolutions of the impeller, were required to achieve a steady, cyclical variation of the ow variables. The power number was then calculated from the pressure distribution on the impeller. Mixing time calculations were then initiated with this velocity eld as the starting point, i.e., for a case where the ow eld is already welldeveloped. The initial tracer concentration was set to zero throughout the ow domain except in a small volume (typically about 0.01% of the tank volume) where mass fraction of tracer was set to 0.5. (The location of the tracer injection may have an effect on the overall mixing time calculated. Preliminary calculations have shown this variation to be of the order of 10% in a well-mixed tank. While this is small in comparison with the overall variation of mixing time with Reynolds number, in order to minimize this effect, the tracer injection point was kept the same in all the simulations.) The evolution of the concentration eld with the introduction of tracer was then calculated by marching forward in time. Mixing time was obtained when the tracer mass fraction lay within a speci ed interval throughout the vessel. Since the simulated ow domain corresponds to only one-eighth of the tank, the simulated dye tracing experiment corresponds to the case where the tracer is added at eight points simultaneously in the tank. Since this is not the case usually in mixing time studies, the tracer injection point was chosen to lie close to the shaft near the impeller. This would correspond to a case where the tracer is added radially into the tank through the impeller shaft and is thus more realistic. While the location of the tracer injection is expected to make some difference to the mixing time, the effect has not been studied here. RESULTS AND DISCUSSION Flow Field The typical velocity eld produced for a laminar ow case, a transitional ow case and for a turbulent ow case are shown in Figure 2(a), 2(b) and 2(c), respectively, along a radial plane in the plane of the impeller. Here, the laminar ow case corresponds to an impeller Reynolds number (NRe) of 0.96, the transitional ow case to an NRe of 96 and the turbulent ow case to an NRe of 960. The geometry of the impeller and the vessel, speed of agitation are

Figure 2. Vector plot of the predicted ow eld in a radial plane for the case of: (a) laminar ow at an impeller Reynolds of 0.96; (b) for transitional ow at a Reynolds number of 96; (c) for turbulent ow at an impeller Reynolds number of 960 using: (i) the standard (high Reynolds number) k-e model; (ii) using the low Reynolds number k-e model. The geometrical details and the speed of rotation are the same in all cases, only the viscosity of the uid is changed to modify the Reynolds number.

identical in all the three cases; the kinematic viscosity in the laminar, transitional and turbulent ow cases is 10 3, 105 and 106 m2 s1, respectively. The turbulent ow case corresponds to that investigated experimentally1 and comparison with the data is shown later. Examination of the ow patterns produced in Figure 2 shows that for Trans IChemE, Vol 80, Part A, July 2002

CFD STUDY OF POWER AND MIXING TIME laminar ow (Figure 2(a)), a small circulation loop is set up and signi cant velocities are created only close to the impeller. The region away from the impeller appears to be a dead zone. At the higher Reynolds number of 96 (see Figure 2(b)), a stronger circulation pattern extending over a larger volume of the vessel is created. A low velocity region persists away from the shaft at the top of the vessel. In the last case corresponding to a Reynolds number of 960, the velocity eld calculated using the standard (high Reynolds number) k-e turbulence model and the low Reynolds number k-e turbulence model are shown in Figure 2(c(i)) and 2(c(ii)), respectively. The standard k-e model (see Figure 2(c(i))) predicts a strong circulation pattern which affects every part of the vessel and there are no dead zones. The low Reynolds number k-e model (see Figure 2(c(ii))) prediction shows considerable deviation from the predicted ow eld obtained using the standard k-e model in the top part of the vessel. The predicted ow pattern here resembles more closely the pattern shown in Figure 2(b) for laminar ow at a Reynolds number of 96. Indeed, the predicted turbulence level in the low Reynolds number k-e model for NRe of 960 was so low that the solution in Figure 2(c(ii)) corresponds to that of laminar ow. In all cases, the impeller creates a radially outward ow which is de ected from the vessel wall to create a two-cell circulatory pattern in the vessel. A comparison of the predicted velocity variation in the turbulent ow case with the measured data of Dong et al.1 is shown in Figure 3. Here, the radial pro les of the circumferential velocity at different heights in the vessel are shown. The predictions with three cases are compared with the data: (i) those obtained with the standard k-e turbulence model with the same grid as used by Dong et al.4; (ii) those obtained with the same turbulence model but with a grid which is twice as big in each direction; (iii) predictions obtained using the low Reynolds number k-e model. Close to the impeller (see Figure 3(b)), all the three models predict nearly the same velocity pro le which agrees fairly well with data. Away from the impeller, especially in the top half of the vessel (see Figures 3(c)3(e)), there is considerable deviation. The velocity pro les obtained with the low density grid show the same variation as obtained by Dong et al.4 which is not surprising as they also used the same turbulence model. The effect of increasing the grid size is only small and is not enough to make up for the considerable deviation between the predicted and measured velocity data in the top region. However, the velocity pro les corresponding to the low Reynolds number k-e model show very good agreement with the data in this region. Thus, the lack of agreement between the data and the predictions found in Dong et al.s4 simulations appears to be due to the use of the standard k-e turbulence model which is accurate only at high Reynolds numbers. The predicted radial and axial velocity pro les are compared with the experimental data in Figures 4 and 5, respectively. Generally speaking, the radial velocities are nearly an order of magnitude smaller than the circumferential velocity except in the impeller plane. The maximum radial velocity occurs close to the impeller tip and was measured1 to be about 0.5 Wtip. This is in line with the predictions. However, the low Reynolds number k-e model appears to predict a ow pattern which is different from that predicted by the standard k-e models and it is the latter Trans IChemE, Vol 80, Part A, July 2002

487

which appear to be in agreement with the trend exhibited by the data as far as radial and axial velocities are concerned. The turbulent kinetic energy obtained using the standard k-e model is compared with the experimental data in Figure 6. It can be seen that fairly good agreement is obtained between the predictions and the data in the region close to the impeller. However, though the predicted values are signi cantly lower in the top half of the vessel compared to the values prevailing near the impeller, they are still considerably higher than zero (which is what the measured data indicate) and would contribute signi cantly to the effective viscosity. As will be seen later, this has a signi cant effect on the predicted mixing time. The low Reynolds number k-e model predicts a completely laminar ow for this case and hence the predicted turbulent kinetic energy is zero. This is not in agreement with data. It is only at higher Reynolds numbers that it predicts non-zero turbulence level. This case illustrates the dif culty of predicting ows in transitional ow conditions. The use of the standard k-e model predicts the correct level of turbulence near the impeller region but also a relatively high level of turbulence in the whole of the vessel while the low Reynolds number k-e model predicts no turbulence at all throughout the vessel. The data lie in between these two extremes showing a signi cant level of turbulence only close to the impeller. Accordingly, the circumferential velocity pro les agree better with the high Reynolds number turbulence model predictions in the near-impeller region (see Figure 3(b)) and with the low Reynolds number model in the region well away from the impeller (see Figures 3(c)3(e)). However, the incorrect prediction of turbulent kinetic energy, and hence, the effective viscosity by the latter leads to a poor prediction of the circulation pattern in the axial plane resulting in rather poor prediction of the radial and axial velocity pro les. Better turbulence models should be used in this low Reynolds number case and studies of the kind reported by Jenne and Reuss15 assessing various low Reynolds number turbulence models should be carried out with a wide range of parameters. Lack of data, especially of turbulence quantities, precludes such a study at this stage. Power Consumption Laminar ow The power consumption, expressed in terms of the product (NP NRe) where NP is the power number de ned as P=rN 3 d 5 ; and NRe is the impeller Reynolds number, is summarized in Table 1. Here, the predicted (NP NRe) value obtained using the low and the high density grids are compared with the empirical correlations of OConnell and Mack30, Nagata31 and Stein32. It is generally found that this product remains constant in laminar ow. However, the value of the constant varies from case to case. Based on their experimental studies, OConnell and Mack produced the following empirical correlation for straight blade paddle impeller in a vessel with three baf es for laminar ow conditions: NP NRe 90 S 0:327b=D0:635S
0:280

where S is the number of blades, b is the blade width and D is the tank diameter.

488

MURTHY SHEKHAR and JAYANTI

Figure 3. Comparison of the predicted pro les of circumferential velocity with the data of Dong et al.1 at a Reynolds number of 960 at a height of: (i) 0.01 m; (ii) 0.03 m; (iii) 0.05 m; (iv) 0.07 m; (v) 0.09 m, from the bottom of the vessel. The broken line corresponds to results obtained with the low density grid, the thick solid line to the high density grid with standard k-e model and the thin solid line to the low Reynolds number k-e model. The lled diamonds correspond to the measurements of Dong et al.1 The velocity is non-dimensionalized by dividing by the impeller tip velocity.

Trans IChemE, Vol 80, Part A, July 2002

CFD STUDY OF POWER AND MIXING TIME

489

Figure 4. Same as in Figure 3 except that the comparison is made for the radial velocity pro les.

Trans IChemE, Vol 80, Part A, July 2002

490

MURTHY SHEKHAR and JAYANTI

Figure 5. Same as in Figure 3 except that the comparison is made for the axial velocity pro les.

Trans IChemE, Vol 80, Part A, July 2002

CFD STUDY OF POWER AND MIXING TIME

491

Figure 6. Comparison of the predicted turbulent kinetic energy with data of Dong et al.1. Predictions of the low grid values (dotted line) and the high grid values (solid line) and the data (diamonds) are rendered dimensionless by dividing by the square of the tip velocity.

Trans IChemE, Vol 80, Part A, July 2002

492

MURTHY SHEKHAR and JAYANTI


Table 1. Predicted NP NRe values for laminar ow cases. Impeller Reynolds number 0.48 0.96 1.92 4.8 9.6 Predicted (NP NRe) from CFD 81.7 81.5 82.7 86.9 102.9 Predicted (NP NRe) from correlations OConnell30 130.2 130.2 130.2 130.2 130.2 Nagata31 122.7 122.7 122.7 122.7 122.7 Stein32 88.0 88.0 88.0 88.0 88.0

RPM 50 100 200 500 1000

values of NRe although it becomes independent of Reynolds number in the asymptotic limit: 0:66 p A 103 1:2NRe NP B 10 0:66 NRe 103 3:2NRe where A is given by equation (8) above and B and p are given by: B 10f1:34b=D0:5 1:14d=Dg p 1:1 4b=D 2:5d =D 0:52 7b=D4
2

11 12

Nagata31 obtained the following correlation for an unbaf ed vessel for a wide range of width (b) and diameter (d) of the blades: Laminar flow : NP A NRe 7

with A 14 b=Df670d =D 0:62 185g

He found that constant A varied between 22.3 and 260 depending on the combination of b=D and d=D where D is the diameter of the vessel. Unfortunately, none of their experimental conditions matched with those used in the experiments of Dong et al.1 and extrapolation, using equation (7), was necessary to obtain the values calculated in Table 1. Nagatas31 correlation is valid for a two-blade impeller whereas the present study used an eight-blade impeller. This difference has been accounted for using the equivalent blade area method suggested in Nagata31. Application of these three correlations to the present case shows that the product NP NRe must have a constant value of 130.2 and 122.7, according to equations (6) and (7), respectively. Stein32 obtained a constant value of 110 for the product (NP NRe) in his experiments for an unbaf ed vessel. Adjusting this for the difference in the number of blades brings the effective Stein32 value to 88. The computed values shown in Table 1 over the range of NRe between 0.48 and 9.6 show that it varies slightly, increasing from 82 to 102 over this range. Considering the uncertainties (in terms of the presence of baf es, range of geometric parameters and Reynolds number) in the correlations, the agreement between the predictions and the correlations can be considered to be good. Turbulent ow In turbulent ow, it is expected that the power number, NP, de ned as above, is independent of the impeller Reynolds number and is a function of the geometrical characteristics only. OConnell and Mack30 obtained the following relation for the power number under turbulent ow conditions for a three-baf e con guration: Np 9:74S 0:495b=D1:33S
0:108

The results obtained for power consumption under turbulent ow conditions are compared in Table 2 with these two correlations. The range of Reynolds number covered in the calculations was between 960 and 48,000. It can be seen that while there is good agreement at low Reynolds numbers, there is increasing disparity between the CFD predictions and the Nagata correlation with the latter showing a pronounced decrease as Reynolds number increases. This may be attributed to the effect of vortex formation (which has not been considered in the present simulations) on NP which has the effect of reducing NP. This is supported by circumstantial evidence from two independent sources. Firstly, Ciofalo et al.8 reported that they were able to obtain the correct trend of NP vs NRe by taking account of the vortex formation in their CFD simulations carried out using the commercial code CFX in the Reynolds number range of 23,000 to 40,000. Secondly, the correlation of OConnell and Mack30, which has been developed for baf ed vessels in which there is no vortex formation, shows a value of NP which is higher than the CFD prediction and, more importantly, does not show any variation with Reynolds number. Transition regime Computations of power have been carried out for intermediate Reynolds numbers of 24, 48, 96, 192 and 480 using the low Reynolds number k-e model. Since the ow is not expected to be fully turbulent, and since the low Reynolds number k-e model is capable of dealing with local relaminarization of turbulence23,24, it was thought that this would enable the computation of the ows under transitional conditions. The computed NP values are compared in Table 3 with the correlation of Nagata31 given by equation (12) above. It can be seen that the computed NP value decreases as the Reynolds number increases which is in

Table 2. Predicted NP values for turbulent ow. Power number (NP) from CFD RPM 50 100 200 500 1000 2000 5000 Impeller Reynolds number 480 960 1920 4800 9600 19,200 48,000 Std k-e model 8.1 7.6 6.8 6.4 6.7 6.6 6.4 Low Re k-e model 6.21 5.2 6.6 7.5 Power number (NP) from correlation of OConnell30 10.8 10.8 10.8 10.8 10.8 10.8 10.8 Nagata31 7.8 6.9 5.9 4.4 3.4 2.6 1.9

Nagata31 obtained the following empirical correlation which contains some Reynolds number in uence for intermediate

Trans IChemE, Vol 80, Part A, July 2002

CFD STUDY OF POWER AND MIXING TIME


Table 3. Predicted NP values for transition regime ow. Impeller Reynolds number 24 48 96 192 Power number (NP) from CFD low Re k-e model 8.4 7.7 7.4 6.2 Power number from Nagata31 correlation 14.5 11.7 10.1 7.8

493

RPM 25 50 100 200

agreement with the trend predicted by the Nagata correlation. However, the decrease is not as sharp as that exhibited in the correlation values. In all these cases, the predicted turbulence level in the vessels was zero throughout. Thus, the power values obtained above correspond effectively to those for laminar ow. At higher rotational speed (corresponding to an impeller Reynolds number of 1920), the low Reynolds k-e model predicted non-zero turbulence level and, at even higher Reynolds numbers, the turbulence levels predicted by this model were nearly the same as those predicted by the standard k-e model. The difference between using a low Reynolds turbulence model and the standard k-e model for the transition regime would be that the latter would have predicted some (non-zero) level of turbulence even for NRe 96: This may not have much impact on the power consumption (the NP value predicted by the low Reynolds turbulence model was about 25% lower for NRe in the range of 500 to 1000), but may have an effect on the mixing time, as will be shown later. A consolidated plot of the variation of the predicted NP with NRe in the three regimes is shown in Figure 7 along with the empirical correlation of Nagata31. The computations cover the range of 1 < NRe < 48;000: For NRe < 10; no turbulence model was used. For 10 < NRe < 500; the values obtained using the low Reynolds number k-e model are plotted while the values obtained using the standard k-e model are plotted for NRe > 500: It is seen that the overall trend is captured quite well except at high Reynolds numbers where vortex formation in the vessel may have caused the discrepancy. Mixing Time Details of mixing process The progress of mixing is speci c to the ow eld which is characterized by a circulation pattern and an effective

diffusivity. In the case of laminar ow, the diffusivity is due to molecular motion and is constant for isothermal ows. For turbulent ows, the molecular diffusivity is augmented by turbulent uctuations and the effective diffusivity, being a strong function of local velocity gradients, becomes a ow property. The progress of mixing in a given ow situation can be visualized by introducing a virtual tracer and monitoring how its concentration (or its mass fraction) changes with time. This is illustrated in Figures 8(a), 8(b) and 8(c) for the cases of laminar, transitional and turbulent ows, respectively considered in Figure 2. The three ows are such that the impeller location, geometry and speed of rotation are identical; only the uid viscosity is changed to bring about the change in Reynolds number and hence the ow regime. The overall mixing time for the laminar ow case is about 1590 s, about 144 s for the transitional ow case and only about 46 s for the turbulent ow case. In view of this vast difference, the mass fraction contours in Figure 8 are not shown at the same time. For laminar ow, the contours are shown at times of 20 s, 100 s, 500 s and 1000 s from the start of the tracer injection. It can be seen that, the tracer slowly spreads radially outwards and then comes round along the wall into the lower and the upper recirculation loops. The dead zone in the upper left hand corner appears to delay the overall mixing of the tracer. A similar picture emerges for the transitional ow case (in Figure 8(b) showing the mass fraction contours are shown at 8, 16, 64, 128 s) although the mixing is on a faster time scale because of higher velocities. In the turbulent ow case (see Figure 8(c)), the strong radial out ow immediately pushes the tracer into the lower and the upper recirculation loops and the presence of strong recirculation throughout the vessels leads to nearly complete mixing within 36 s from the start of the injection (Figure 8(c(iv))). The predicted progress of mixing with the low Reynolds number k-e model for the same Reynolds number as in Figure 8(c) is shown in Figure 9 where the mass fraction contours at the end of 2, 8, 16, 32, 48, 64 and 80 s are shown in dimensionless form. It is seen that the signi cantly different ow pattern in this case (see Figure 2c(ii)) leads to a considerably different mixing pattern. The main impediment to faster mixing appears to be the additional recirculation zone appearing in the upper corner of the vessel. As a result of this, the overall mixing time increases from 36 s for the standard k-e model to 88 s for the low Reynolds number k-e model. Prediction of overall mixing time The progress of mixing can also be represented in a plot of the variation of the minimum and the maximum tracer mass fraction with time. This plot is shown in Figure 10 for four cases: for laminar ow, for transitional ow, for turbulent ow with standard k-e model and for turbulent ow using the low Reynolds number k-e model. In these plots, the mass fractions are made dimensionless by dividing by the mass fraction under a fully mixed condition. In all four cases, there is a rapid decrease in the maximum concentration in the initial stages of mixing. However, this mixing is very localized in the laminar ow case, as evidenced by the fact that the minimum concentration remains zero during the rst 200 s after the injection of the tracer. In the turbulent and the low Reynolds number

Figure 7. Comparison of the predicted variation of the power number, NP, with the impeller Reynolds number, NRe, with that given by Nagata31. The diamonds indicate CFD computations and the continuous line is from Nagata31.

Trans IChemE, Vol 80, Part A, July 2002

494

MURTHY SHEKHAR and JAYANTI

Figure 8. Progression of mixing of a neutrally buoyant tracer for: (a) laminar ow; (b) transitional ow; (c) turbulent ow. Equi-spaced contours of mass fraction of the tracer, divided by the fully-mixed concentration, are shown for each case at times of: (i) 20; (ii) 100; (iii) 500; (iv) 1000 s from the start of mixing for (a); at times of: (i) 8; (ii) 16; (iii) 64; (iv) 128 s for (b); and at times of (i) 2; (ii) 8; (iii) 16; (iv) 32 s (c).

Trans IChemE, Vol 80, Part A, July 2002

CFD STUDY OF POWER AND MIXING TIME

495

Figure 9. Same as for Figure 8(c) but with the low Reynolds number k-e model. The contours of the mass fraction are shown at times of (i) 2; (ii) 8; (iii) 16; (iv) 32; (v) 64; (vi) 80 s from the start of the tracer injection.

Trans IChemE, Vol 80, Part A, July 2002

496

MURTHY SHEKHAR and JAYANTI

Figure 10. Variation with time of the minimum and the maximum mass fraction in the vessel for the case of: (a) laminar ow at a Reynolds number of 0.96; (b) for transitional ow at a Reynolds number of 96, and for turbulent ow at a Reynolds number of 960 using (c) the standard k-e turbulence model; (d) the low Reynolds number k-e model.

Trans IChemE, Vol 80, Part A, July 2002

CFD STUDY OF POWER AND MIXING TIME


Table 4. Predicted mixing times for laminar ow case. Impeller Reynolds number 0.96 1.92 4.8 9.6 Predicted mixing time (s) using CFD 1590.0 913.0 316.0 149.0 Predicted N y using CFD 2650.0 3043.3 2633.3 2483.3

497

RPM 100 200 500 1000

turbulent ow cases, the initial mixing is mostly by convection as the zero concentration stage lasts only about 15 s in both cases. The case of the transitional ow is similar to that of the laminar ow case but on a reduced time scale of operation. The overall mixing time, de ned here as the earliest time for which the mass fraction in the entire vessel remains within 5% of the fully-mixed value, varies from 36 s in case of turbulent ow to about 1590 s in the case of laminar ow. The predicted mixing time values for various Reynolds number in laminar ow are compared in Table 4. Due to the large amount of time required for the computation of the mixing time, it has not been possible, in some cases, to carry out calculations until complete mixing is achieved and the mixing times in these cases have been obtained by extrapolation. In all cases, mixing time calculations were conducted until the minimum concentration in the vessel reached at least 90% of the fully mixed concentration. It can be seen that as the Reynolds number increases, the mixing time decreases, roughly inversely. This result is somewhat surprising as it is thought31 that the product (rotational speed mixing time), i.e., (N y) is expected to decrease with Reynolds number. However, this argument is based on extrapolation of the mixing time values obtained in the transitional regime NRe > 10: Recently, Stein32 reported mixing time data in unbaf ed and baf ed vessels over a range of Reynolds numbers. His results for baf ed vessels clearly show a decrease of (N y) with increasing NRe even in the laminar region. However, the results for unbaf ed vessels (which are much fewer in number in the laminar region) show a trend towards a constant value of (N y) of about 2500 which agrees quite well with the results shown in Table 4. Also, results of mixing studies in unbaf ed vessels using anchor, screw and helical ribbon impellers33,34 also show that (N y) is constant for NRe < 10: This is also supported by the data of Stein32 for a screw impeller. Considering all these points, it may be concluded that

even for paddle impellers, (N y) approaches a constant value in unbaf ed vessels for very low Reynolds numbers. The predicted mixing time values for various Reynolds number in the transitional and in the turbulent ow regimes are compared in Table 5. Here, the predicted mixing times using the k-e model and the low Reynolds number k-e turbulence model are shown at various Reynolds numbers 24 to 4800. Also shown are the mixing times obtained using the empirical correlations of Nagata31 which are applicable for unbaf ed vessels with paddle impellers and are valid in the Reynolds number range under investigation as the values were obtained from a graph drawn as a function of Reynolds number. It can be seen that the (N y) value decreases slightly in the transitional regime but that it remains more or less constant at higher Reynolds numbers. In the turbulent ow regime, the results obtained with the low Reynolds k-e model match very well with those of Nagata while the results from the standard k-e model are much lower. This shows that although the prediction of power is not affected signi cantly by which grid and which turbulence model is used, the mixing time is in uenced by the choice of these parameters. CONCLUSIONS In the present paper, the ow eld, power consumption and mixing time were predicted using the commercial CFD code CFX for an eight blade paddle impeller in an unbaf ed vessel over a range of impeller Reynolds numbers covering laminar, transitional and turbulent regimes. The predicted velocity eld in one case (corresponding to a Reynolds number of 960) were compared with the experimental data of Dong et al.1. This showed while the velocity eld and the turbulence level predicted by the standard k-e model showed good agreement with data in the near-impeller region, the low Reynolds number k-e model gave a better result for the velocity eld away from the impeller. This indicates that at this Reynolds number, the ow is not fully turbulent. Comparison of the predicted power and mixing time using the two turbulence models showed that while the difference in power prediction was only of the order of 25%, the mixing time varied by more than 100%. Comparison with the correlations of Nagata31 showed that the prediction of the low Reynolds k-e model is likely to be the more accurate. Calculations in the laminar regime showed that 1 NP NRe which is consistent with experimental data31,32. However, the mixing time calculations showed that the

Table 5. Predicted mixing times for transitional and turbulent ow. Impeller Reynolds number 24 48 96 480 960 1920 4800 Predicted mixing time std. k-e model 78.0 36.0 23.0 13.0 low Re k-e model 833.0 412.5 144.0 173.4 88.0 36.0 19.0 Nagata31 correlation 960.0 384.0 168.0 158.0 73.0 36.5 14.6 std. k-e model 65.0 76.7 76.7 108.3 Predicted N y low Re k-e model 347.1 343.8 240.0 144.5 146.7 120.0 158.3 Nagata31 correlation 400.0 320.0 280.0 131.7 121.7 121.7 121.7

RPM 25 50 100 50 100 200 500

Trans IChemE, Vol 80, Part A, July 2002

498

MURTHY SHEKHAR and JAYANTI


unbaf ed mixing vessel provided with a pitched-blade turbine, IChemE Symp Ser, 136: 349356. Jenne, M. and Reuss, M., 1999, A critical assessment on the use of k-e turbulence models for simulation of the turbulent liquid ow induced by a Rushton-turbine in baf ed stirred-tank reactors, Chem Eng Sci, 54: 39213941. Chen, Y. S. and Kim, S. W., 1987, Computation of turbulent ows using an extended k-e turbulence closure model, NASA CR-17920. Armenante, M. P., Luo, C., Chou, C. C., Fort, I. and Medek, J., 1997, Velocity pro les in a closed, unbaf ed vessel: Comparison between experimental LDV data and numerical CFD predictions, Chem Eng Sci, 52: 34833492. Tanguy, P. A., Thibault, F. and Brito de la Fuente, E., 1996, A new investigation of the Metzner-Otto concept for anchor mixing impellers, Can J Chem Eng, 74: 222228. De la Villeon, J., Bertrand, F., Tanguy, P. A., Labrie, R., Bousquet, J. and Lebouvier, D., 1998, Numerical investigation of mixing ef ciency of helical ribbons, AIChE J, 44: 972977. Ranade, V. V Bourne, J. R. and Joshi, J. B., 1991, Fluid mechanics and ., blending in agitated tanks, Chem Eng Sci, 46: 18831893. Warsi, Z. U. A., 1993, Fluid Dynamics: Theoretical and Computational Approaches (CRC Press, Louisiana, USA). Rushton, J. H., Costich, E. W. and Everett, H. J., 1950, Power characteristics of mixing impellersPart I, Chem Eng Prog, 46(8): 395404. Jones, W. P. and Launder, B. E., 1972, The prediction of laminarization with a two-equation model of turbulence, Int J Heat Mass Trans, 15: 301314. Jayanti, S. and Hewitt, G. F., 1997, Hydrodynamics and heat transfer in wavy annular gas-liquid ow: A computational uid study, Int J Heat Mass Trans, 40: 24452470. Harvey, A. D., Wood, S. P. and Leng, D. E., 1997, Experimental and computational study of multiple impeller ows, Chem Eng Sci, 52: 14791491. Jayanti, S., 2001, Hydrodynamics of jet mixing in vessels, Chem Eng Sci, 56: 193210. Murthy, J. Y., Mathur, S. R. and Choudhury, D., 1994, CFD simulation of ows in stirred tank reactors using a sliding mesh technique, IChemE Symp Series, 136: 341348. Van Doormal, J. P. and Raithby, G. D., 1984, Enhancements of the SIMPLE method for predicting incompressible uid ows, Numerical Heat Trans, 7: 147163. Rhie, C. M. and Chow, W. L., 1983, Numerical study of the turbulent ow past an airfoil with trailing edge separation, AIAA J, 21: 1527 1532. OConnell, F. P. and Mack, D. E., 1950, Simple turbines in fully baf ed tankspower characteristics, Chem Eng Prog, 46(7): 358368. Nagata, S., 1975, Mixing- Principles and Applications (Kodansha Ltd, Tokyo, Japan). Stein, W. A., 1992, Mixing times in bubble columns and agitated vessels, Int Chem Eng, 32: 449474. Hoogendoorn, C. J. and Den Hartog, A. P., Modern studies on mixers in the viscous ow region, Chem Eng Sci, 22: 16891699. Novak, V. and Rieger, F., Homgoenization ef ciency of helical ribbon and anchor agitators, Chem Eng J, 9: 6370.

product (N y) was constant. This is not consistent with the extrapolations of Nagata31 or with the experimental results of Stein32 for baf ed vessels but is consistent with measurements in unbaf ed vessels for anchor, screw and helical ribbon impellers33,34. The few results of Stein32 for unbaf ed vessels also tend to con rm this trend. Therefore, it may be concluded that the product (N y) is constant for paddle mixing in unbaf ed vessels under laminar ow conditions. The computations in the transitional and turbulent ow regimes indicate that good agreement with data may be obtained by using a turbulence model corrected for low Reynolds number=local laminarization effects in the Reynolds number range of 10 to 1000. At high Reynolds numbers NRe > 10;000; it appears that effects associated with vortex formation and free surface deformation should be taken into account in the simulations in order to get correct results. On the whole, the computations reported in the present paper show that reasonable predictions of the velocity eld, power consumption and mixing time can be obtained without any empiricism, other than that associated with turbulence modelling, being introduced in the simulations. CFD simulations should therefore be readily applicable to investigate scale-up effects in mixing. REFERENCES
1. Dong, L., Johansen, S. T. and Engh, T. A., 1994, Flow induced by an impeller in an unbaf ed tank-II, numerical modelling, Chem Eng Sci, 49: 35113518. 2. Uhl, V W. and Gray, J. B., (eds), 1966, MixingTheory and Practice . (Academic Press, New York, USA). 3. Harnby, N., Edwards, M. F. and Nienow, A. W., 1985, Mixing in the Process Industries (Butterworths, London, UK). 4. Dong, L., Johansen, S. T. and Engh, T. A., 1994, Flow induced by an impeller in an unbaf ed tankI, experimental, Chem Eng Sci, 49: 549 560. 5. Randade, V V and Dommeti, S. M. S., 1996, Computational snapshot . . of ow generated by axial impellers in baf ed stirred vessels, Trans IChemE, Part A, Chem Eng Res Des, 74(A7): 576484. 6. Bakker, A., Myers, K. J., Ward, R. W. and Lee, C. K., 1996, The laminar and turbulent ow pattern of a pitched blade turbine, Trans IChemE, Part A, Chem Eng Res Des, 74(A) 485491. 7. Xu, Y. and McGrath, G., 1996, CFD predictions of stirred tanks ows, Trans IChemE, Part A, Chem Eng Res Des, 74(A6): 471475. 8. Ciafalo, M., Brucato, A., Grisa , F. and Torraca, N., 1996, Turbulent ow in closed and free surface unbaf ed tanks stirred by radial impellers, Chem Eng Sci, 51: 35573573. 9. Bakker, A., Laroche, R. D., Wang, M. H. and Calabrese, R. V 1997, ., Sliding Mesh simulation of laminar ow in stirred reactors, Trans IChemE, Part A, Chem Eng Res Des, 75(A1) 4244. 10. Naude, I., Xuereb, C. and Bertrand, J., 1998, Direct prediction of the ows induced by a propeller in an agitated vessel using an unstructured mesh, Can J of Chem Eng, 76: 631639. 11. Ng, K., Fentiman, N. J., Lee, K. C. and Yianneskis, M., 1998, Assessment of sliding mesh CFD predictions and LDA measurements of the ow in a tank stirred by a Rushton impeller, Trans IChemE, Part A, Chem Eng Res Des, 76(A8): 737747. 12. Launder, B. E., 1989, The prediction of force eld effects on turbulent shear ows via second-moment closure, Advances in Turbulence 2, Fernholz, H. H. and Fiedler, H. E., (eds) (Springer-Verlag, Berlin, Germany). 13. Crochet, M. J., Davies, A. R. and Walters, K., 1984, Numerical Simulation of Non-Newtonian Flow (Elservier, Amsteradam, The Netherlands). 14. Armenante, P., Chou, C. C. and Hemrajani, R. R., 1994, Comparison of experimental and numerical uid velocity distribution pro les in an

15.

16. 17.

18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. 31. 32. 33. 34.

ACKNOWLEDGEMENTS
The computations described here have been performed at the CFD Centre, IIT-Madras, India.

ADDRESS
Correspondence concerning this paper should be addressed to Dr S. Jayanti, Department of Chemical Engineering, IIT-Madras, Chennai 600 036, India. E-mail: sjayanti@iitm.ac.in The manuscript was received 10 May 2001 and accepted for publication after revision 26 March 2002.

Trans IChemE, Vol 80, Part A, July 2002

Das könnte Ihnen auch gefallen