Abstract
An open-source CFD software OpenFOAM® is used to simulate two multiphase stirred-tank reactors relevant to industrial processes such as slurry polymerization and fuel production. Gas-liquid simulations are first performed in a single-impeller stirred-tank reactor, studied experimentally by Ford, J. J., T. J. Heindel, T. C. Jensen, and J. B. Drake. 2008. “X-Ray Computed Tomography of a Gas-Sparged Stirred-Tank Reactor.” Chemical Engineering Science 63: 2075–85. Three impeller rotation speeds (200, 350 and 700 rpm) with three different bubble diameters (0.5, 1.5 and 2.5 mm) are investigated. Flow patterns compared qualitatively to those from experiments. Compared to the experimental data, the simulations are in relatively good agreement for gas holdup in the reactor. The second multiphase system is a multi-impeller stirred-tank reactor, studied experimentally by Shewale, S. D., and A. B. Pandit. 2006. “Studies in Multiple Impeller Agitated Gas-Liquid Contractors.” Chemical Engineering Science 61: 486–504. Gas-liquid simulations are performed at two impeller rotation speeds (3.75 and 5.08 RPS). The simulated flow patterns agree with published pictures from the experiments. Gas-liquid-solid simulations of the multi-impeller stirred-tank reactor are also carried out at impeller rotation speed 5.08 RPS. The addition of solid particles with a volume fraction characteristic of slurry reactors changes the flow pattern significantly. The bottom Rushton turbine becomes flooded, while the upper pitched-blade downflow turbines present a radial-pumping flow pattern instead of down-pumping. Nonetheless, the solid phase has a similar flow pattern to the liquid phase, indicating that the particles modify the effective density of the fluid.
1 Introduction
Stirred-tank reactors (STRs) are widely used in industrial processes, such as in the chemical, biological, pharmaceutical, as well as many other industries. Usually, multiphase systems are of most interest in these processes, including two-phase (gas-liquid/gas-solid/solid-liquid/liquid-liquid) systems and three-phase (gas-liquid-solid) systems (Achouri et al. 2012; You et al. 2014). Gas-liquid STRs with a single impeller have been studied extensively using experiments. Investigations of the effects of reactor parameters and operating conditions such as the geometry of the tank, impeller type, size and location, impeller rotation speed and gas flow rate on the reactor characteristics including flow regimes and patterns, gas hold-up, power-consumption, mixing time and mass-transfer coefficient between the phases have been carried out for decades (Ford et al. 2008; Joshi et al. 2011a; Kumaresan and Joshi 2006; Miller 1974; Nienow et al. 1985; Nienow 1998; Parasu-Veera, Patwardhan, and Joshi 2001; Rewatkar et al. 1993; Sawant and Joshi 1979). For large-scale, tall-thin, industrial gas-liquid STRs, multiple impellers are usually preferred over a single impeller because multiple impellers can have better gas utilization due to the high residence time of gas bubbles, maintain homogeneity in the reactor, provide high surface-to-volume ratio and offer lower shear than single-impeller STRs for shear-sensitive systems (Ahmed et al. 2010; Dutta and Pangarkar 1995; Himmelsbach et al. 2006; Kasat and Pandit 2004; Khopkar et al. 2006; Shewale and Pandit 2006; You et al. 2014). However, the complexity of the flow in STRs with multiple impellers increases with increasing number of impellers as any change in the reactor parameters, such as tank aspect ratio, number, type, size, location and configuration of impellers and operating conditions, may influence the reactor operating characteristics significantly. Therefore, many researchers have studied multiple-impeller STRs experimentally (Kasat and Pandit 2004; Shewale and Pandit 2006; Vr’abel et al., 1999, 2000; You et al. 2014) for gas-liquid two-phase flows. On the other hand, experiments of gas-liquid-solid STRs with single or multiple impellers are reported less in the literature than gas-liquid STRs (Boyer, Duquennt, and Wild 2002; Chen, Bao, and Gao 2009; Conway, Kyle, and Rielly 2002; Dohi et al., 2002, 2004; Dutta and Pangarkar 1995; Fishwick, Winterbottom, and Stitt 2003; Joshi et al. 2011a; Micale et al. 2000; Nienow and Bujalski 2002; Rewatkar and Joshi 1992), as the addition of a solid phase increases the flow complexity in the reactor and greatly limits the range of optical-based measurement techniques that can be applied.
Computational fluid dynamics (CFD) simulation is another approach to study mixing and fluid dynamics in STRs. While, many single-phase CFD simulations of STRs with single or multiple impellers has been reported in literature, for gas-liquid systems the main focus is devoted to single impeller STRs (Achouri et al. 2012; Arlov, Revstedt, and Fuchs 2008; Bakker and Van Den Akker 1994; Gosman et al. 1992; Jahoda, Tomaskove, and Moste 2009; Jahoda et al. 2007; Joshi et al. 2011a; K’alal, Jahoda, and Fort 2014; Lane, Schwarz, and Evans 2005; Luo, Issa, and Gosman 1994; Moilanen et al. 2008; Murthy and Joshi 2008; Petitti et al. 2013; Ranade and Van Den Akker 1994; Scargiali et al. 2007; Tyagi et al. 2007) and in these studies, the effects of reactor parameters and operating conditions, and different models such as turbulence models, drag correlations and bubble coalescence and breakage models have been studied. As mentioned above, due to complicated flow regimes and patterns in gas-liquid STRs with multiple impellers, not as many works as for single-impeller STRs are reported in the literature (Ahmed et al. 2010; Joshi et al. 2011b; Kerdouss, Bannaari, and Proulx 2006; Khopkar and Tanguy 2008; Khopkar et al. 2006; Min et al. 2008; Tyagi et al. 2007). Even more, for gas-liquid-solid systems, very few studies are reported for single-impeller STRs (Murthy, Ghadge, and Joshi 2007; Panneerselvam, Savithri, and Surender 2009) while for multi-impeller STRs, works found in the literature are very limited and almost all of these studies are performed using commercial CFD software, such as ANSYS FLUENT® and CFX®.
In this work, an open-source CFD software, OpenFOAM®, is used to numerically investigate two stirred-tank reactors studied experimentally (Ford et al. 2008; Shewale and Pandit 2006), but also of interest for industrial applications such as slurry reactors. Gas-liquid simulations are first performed in a single-impeller STR, reported in (Ford et al. 2008). Then a multi-impeller STR, studied experimentally by (Shewale and Pandit 2006), is simulated for both gas-liquid and gas-liquid-solid systems. However, it should be noted, in their study, Shewale and Pandit (2006) only studied gas-liquid two phase system. On the other hand, in this study, in order to observe the effect of the addition of solid phase to the gas and liquid phase individually, three-phase (gas-liquid-solid) simulations are also presented. In the remainder of this work, Section 2 describes modeling of the multiphase system. The equations used to model the system are outlined. Section 3 introduces the two simulated STRs, and the simulation conditions are described in detail. Section 4 discusses the simulation results, where Section 4.1 presents results obtained in the single-impeller STR and comparison to experiments, and Section 4.2 presents gas-liquid simulation results of the multi-impeller STR. Gas-liquid-solid simulations are discussed in Section 4.3. Conclusions are drawn in Section 5. Results for parameters outside the range reported in the main text are available in the Supplementary Materials.
2 Modeling of the multiphase system
2.1 Governing equations
The Eulerian multi-fluid method (Drew 1983) is used to model the systems in this work. All phases are treated as interpenetrating continua and represented by their volume fractions. The governing equations are summarized below.
The continuity equations are written as
where
The momentum equation is written as
where p is the shared pressure, τk is the stress tensor of phase k, g is the acceleration due to gravity and Mjk represents the momentum exchange between phases j and k. The momentum exchange between phases can include contributions from the following forces: drag, lift, virtual mass, turbulent dispersion and wall lubrication. In this work, drag, virtual mass and turbulent-dispersion forces are included. The buoyancy force results in the shared-pressure term in Eq. (2). The turbulent-dispersion force is required to ensure that the multi-fluid model is hyperbolic when a separate bubble pressure is not employed. Due to the intense mixing provided by the impellers, other exchange forces such as lift were neglected.
2.1.1 Interphase drag correlation
The drag force between the gas and liquid, or the solid and liquid phases is expressed as
where the dispersed phase can be the gas (k = g) or solid phase (k = s), and the coefficient Kkl depends on the specific drag correlation.
The drag correlation of (Tomiyama et al. 1998) is used for the drag force between the gas and liquid phases:
where dg is the diameter of gas bubbles, and the drag coefficient CD is
The bubble Reynolds number Re and Eötvös number Eo are defined as
with
The Wen and Yu drag correlation (Wen and Hu 1966) is used for fluid-solid interactions:
where
and the slip Reynolds number Re is
with ds being the representative particle diameter (taken to be constant).
2.1.2 Virtual mass force
A constant virtual-mass coefficient (CVM = 0.5) (Drew and Lahey 1987) is used to account for the gas-liquid virtual mass force, defined by
where
2.1.3 Turbulent-dispersion force
The gas-liquid turbulent-dispersion force can be expressed as
where a constant turbulent-dispersion coefficient (CTD = 1.0) (Burns et al. 2004) is used, and kl is the turbulent kinetic energy of the liquid. This force is required to keep the multi-fluid model hyperbolic when the bubble pressure is neglected in the shared-pressure model.
In summary, in Eq. (2),
2.2 Turbulence models
In the present work, large-eddy simulations (LES) are used to account for the turbulence in the reactor for each phase. The sub-grid scale (SGS) viscosity for phase k is calculated using the Smagorinsky sub-grid model (Smagorinsky 1963):
where the Smagorinsky constant CS = 0.168 in this work. Δ is the filter width calculated as the cubic root of the volume of each cell, and Dk is the rate-of-strain tensor of phase k. The stress tensor of phase k, τk in Eq. (2), is expressed as
where μk, lam is the dynamic viscosity of phase k, and I is the unit tensor. The SGS turbulent kinetic energy of phase k can be found by assuming local equilibrium for kk, expressed as
where the SGS stress tensor
2.3 Multiple reference frame (MRF) model
The rotation of the impellers is treated by using the MRF model (Luo, Issa, and Gosman 1994), which divides the computational region into a rotating region and a stationary region. In the stationary region, governing equations are solved in a rotating reference frame. At the interface of the two regions, a reference frame transformation is performed.
The fluid velocities can be transformed from the stationary reference frame to the rotating reference frame by calculating the relative velocities as
where Ur, k is the relative velocity of phase k viewed from the rotating reference frame, Uk is the absolute velocity viewed from the stationary reference frame, ω is the angular velocity of the rotating coordinate system relative to the stationary reference frame, and r is the position vector of a point in the rotating region with respect to the origin of the rotating reference frame. When the governing equations are solved in the rotating reference frame, the continuity equation is written as
The momentum equation in the rotating reference frame is expressed as
where the extra term
3 Simulated reactors and operating conditions
Two stirred-tank reactors are studied in the present work. The first reactor was investigated experimentally by (Ford et al. 2008) and is referred to as Reactor 1. The second reactor, referred to as Reactor 2, was investigated experimentally by (Shewale and Pandit 2006). Details about these reactors and the operating conditions are described next.
3.1 Description of reactor 1 and operating conditions
Reactor 1, as described in Ford et al. (2008), is a cylindrical vessel with a flat bottom and four baffles. A six-blade Rushton turbine (RT) is mounted on the shaft near the bottom of the reactor. Air at 300 K is injected uniformly from a ring sparger located at the bottom of the tank into water (300 K). Details on the tank and impeller geometries are listed in Table 1.
Geometric parameters for reactor 1.
Parameter | Value (m) |
---|---|
Tank height, H | 0.300 |
Tank diameter, T | 0.210 |
Baffle width | 0.018 |
Baffle thickness | 0.006 |
Impeller diameter, D | 0.076 |
RT distance to bottom | 0.057 |
RT blade height | 0.019 |
RT blade thickness | 0.003 |
Impeller hub diameter | 0.031 |
Ring sparger diameter | 0.051 |
The commercial mesh generation software Pointwise® is used to generate 420,624 three-dimensional hexahedral and 2436 polyhedral cells with 1,308,248 faces for Reactor 1, as shown in Figure 1. The grid convergence study is performed and given in Table S1 in Supplementary Materials. The gas-liquid flow is found from the two-fluid solver reactingTwoPhaseEulerFoam in OpenFOAM® (OpenFOAM 2015). The initial water level in the reactor is 0.21 m. Air is injected uniformly from the sparger at gas flow rate 1.5 × 10−4 m3/s (9 LPM in Ford et al. 2008). Three impeller rotation speeds are considered 200, 350 and 700 rpm. No-slip boundary conditions are used for both the gas and liquid phases at the tank wall and baffles. Constant gas inlet velocity and atmospheric pressure are used at the gas inlet and outlet, respectively. At the shaft and impeller surfaces, zero relative velocity (Ur, k = 0) is assumed for both phases. The simulated flow time for all cases of Reactor 1 is 65 s, and results are time averaged over the last 60 s.

(a) Illustration of the computational mesh for reactor 1. (b) Enlarged view of the mesh near the impeller.
3.2 Description of reactor 2 and operating conditions
The stirred-tank reactor, studied by (Shewale and Pandit 2006) and referred to as Reactor 2, is a cylindrical vessel with a flat bottom and four baffles. Three impellers are mounted on the shaft and rotate synchronously. The bottom impeller is a six-blade RT, and the middle and top impellers are pitched-blade downflow turbines (PBTD). Details on the tank and impeller geometries are listed in Table 2. In total 669,888 three-dimensional hexahedral cells are generated for Reactor 2 using Pointwise®, as shown in Figure 2. The grid convergence study for Reactor 2 is given in Table S2 in Supplementary Materials.
Geometric parameters for reactor 2.
Parameter | Value (m) |
---|---|
Tank height, H | 1.000 |
Tank diameter, T | 0.300 |
Baffle width | 0.030 |
Ring sparger diameter | 0.100 |
Sparger distance to bottom | 0.075 |
Impeller diameter, D | 0.100 |
Impeller spacing, S | 0.300 |
RT distance to bottom | 0.150 |
RT blade height | 0.020 |
RT blade length | 0.025 |
PBTD height | 0.030 |
PBTD blade angle | 45° |

(a) Illustration of the computational mesh for reactor 2. (b, c) Enlarged views of the mesh near the RT impeller (bottom) and PBTD impeller (top).
The OpenFOAM® solver reactingTwoPhaseEulerFoam (OpenFOAM 2015) is used to solve for the gas-liquid (air-water) flow in the reactor. Initially the water level in the tank is 0.87 m. A gas-liquid-solid flow is also simulated with OpenFOAM® solver reactingMultiphaseEulerFoam (OpenFOAM 2015). The three phases are air, water and polymethyl methacrylate (PMMA) particles, respectively. The initial mixture level is 0.81 m, with solid and liquid volume fractions 0.2 and 0.8, respectively. For both two-phase and three-phase systems, air at 300 K is injected uniformly from the sparger located below the RT into a water-PMMA mixture at 300 K with superficial velocity 0.01 m/s. A uniform bubble diameter (0.5 cm) is used for all simulations of Reactor 2. For the gas-liquid-solid simulations, the particle diameter is assumed constant at 150 µm with density 1190 kg/m3. The impellers are rotating at speeds of 3.75 and 5.08 rps in the down-pumping direction for the gas-liquid simulations, and a rotation speed of 5.08 rps for the gas-liquid-solid simulation. At the tank wall and baffles, a no-slip boundary condition is used for all phases. Constant gas inlet velocity and atmospheric pressure are used at the gas inlet and outlet, respectively. At the shaft and impeller surfaces, zero relative velocity (Ur, k = 0) is used for all phases. Simulated flow time for the gas-liquid system is 65 s, and results are time averaged over the last 60 s. For the gas-liquid-solid system, results are time averaged over 40 s due to the heavy computational cost.
4 Results and discussion
In (Ford et al. 2008), X-ray computed tomography (CT) images of the local gas holdup obtained at different planes are presented. They also report local gas holdup along a line at a specific height and average gas holdup at different heights. Section 4.1 discusses simulation results of this work and compares simulation results to the experimental data. In Shewale and Pandit (2006), experimental results obtained for a gas-liquid flow are discussed, including flow patterns for different regimes and the relationship between the overall gas holdup and power consumption. Section 4.2 presents the gas-liquid simulation results of Reactor 2 and comparisons to experimental data. For the gas-liquid-solid system, no experiment results are reported in the literature, but this system is of interest for slurry polymerization and other industrial processes. Section 4.3 therefore discusses simulation results only.
4.1 Gas-liquid simulations of reactor 1
The overall gas holdup in the vessel
where H0 is the initial water level in the tank (0.21 m). Table 3 lists the averaged mixture height and overall gas holdup in the tank at different impeller rotation speeds and for three bubble diameters of 0.5, 1.5 and 2.5 mm. The comparison of other bubble diameters (3.5 and 5 mm) with the experimental gas holdups are given in the Supplementary Materials (Table S3). The simulation results are compared to experimental data reported in Ford et al. 2008, where
Overall gas holdup in the vessel for three bubble sizes and experiments.
2.5 mm | 1.5 mm | 0.5 mm | Ford et al. | |||||
---|---|---|---|---|---|---|---|---|
Rotation speed | H (m) | αoverall, % | H (m) | αoverall, % | H (m) | αoverall, % | αCT, % | αglobal, % |
200 | 0.215 | 2.3 | 0.216 | 2.9 | 0.216 | 2.9 | 3 | 2.9 |
350 | 0.216 | 2.9 | 0.217 | 3.2 | 0.218 | 3.4 | 3.5 | 3.3 |
700 | 0.223 | 5.6 | 0.225 | 6.5 | 0.225 | 6.7 | 6.9 | 6 |
Figure 3 compares the experimental data for the gas holdup to the simulation results as a function of the height and radial position in the reactor for three bubble sizes. In Figure 3 (a, c and e) the gas volume fractions along a line (x-axis in Ford et al. 2008) at height 0.065 m is plotted for bubble diameters of 0.5, 1.5 and 2.5 mm. From the experimental results the gas volume fraction increases in the impeller region. Although a similar trend is observed from simulations, in general, the simulations underestimate the gas volume fraction, and only the 0.5 mm bubble diameter at 700 rpm mixing rate agrees well with experiments. On the other hand, as shown in Table 3 the average gas hold-up values are well correlated with the experimental results where

Comparison of simulated and experimental gas volume fractions as a function of height and radial position. Black lines with circles, red lines with triangles and blue lines with rectangles represent the experimental results for 700, 350 and 200 rpm mixing rates respectively, and corresponding colored symbols represent simulation results: (a, c, e) Gas volume fraction along a line at height 0.065 m for 0.5, 1.5 and 2.5 mm bubble diameter. (b, d, f) Averaged gas volume fractions at different heights for 0.1, 1.5 and 2.5 mm bubble diameter at different heights.
These results are more obvious in Figure 3(b, d, f) which provides the comparison of gas volume fractions along the reactor in the axial direction. As given in Table 1, the distance of the end of blades from the bottom is 0.076 m and up to this height the experimental and simulation results are different yet in closeness, proximity. Away from the impeller blades, the simulation results are almost the same as the experimental measurements. In comparison of Figure 3(a, b) at high mixing rate (700 rpm – dispersed regime) although the gas holdup is estimated well in the impeller region with 0.5 mm bubble diameter, better agreement is obtained for higher bubble diameters, 1.5 and 2.5 mm, outside of this region. Therefore, it can be concluded that the bubble size distribution is most likely not uniform in the reactor and break-up of bubbles happens within the impeller region due to applied shear and coalescence is the dominant mechanism above the blades. Furthermore, in the loaded regime (350 rpm) good agreement is obtained with 1.5 mm bubble diameter (Figure 3(d)) since a more uniform velocity profile (hence shear applied on the bubbles) can be expected in this regime. This conclusion is supported by the gas and liquid phase velocity magnitude profiles shown in Figure 4 (and in Supplementary MaterialsFigure S4).

Gas velocity magnitude profiles for bubble diameter d = 1.5 mm at different mixing rates and different heights (0.065, 0.087 and 0.11 m): (a–c) 200 rpm, (d–f) 350 rpm, (g–i) 700 rpm.
In Figure 4(a–c) contour plots of the gas-phase velocity magnitudes for the bubble diameter of 1.5 mm are given at impeller rotation speed 200 rpm at three heights 0.065, 0.087 and 0.11 m, corresponding to z = 0.8, 3.0 and 5.3 cm in Figure 5 of (Ford et al. 2008), respectively. (The gas-phase velocity profiles for d = 0.5 and 2.5 mm are given in Supplementary MaterialsFigures S1 and S2). Comparing the three different bubble diameters, although the velocity profiles are similar, the velocity magnitudes increase proportionally with the bubble size. These results clearly indicate that the impeller is flooded at this rotation speed and gas flow rate. The bubbles concentrate near the impeller and shaft and are barely dispersed away from the impeller region (Paglianti 2002) and bubbles spread radially due to the decreased pressure and bubble rise velocity (Ford et al. 2008). Compared to (Ford et al. 2008), similar patterns are observed, i.e., gas holdup is high near the blades, and bubbles spread radially with increasing vertical distance. The gas volume fraction is very low near the tank wall. (In the Supplementary Materials, the flow pattern of water at impeller rotation speed 200 rpm is shown in Figure S4(a–c) at the same heights with the gas phase velocities. The flow pattern of water for 0.5 and 2.5 mm bubble diameters are also provided in Figures S3 and S5). For all bubble diameters no radial-pumping flow pattern can be observed because the impeller is flooded.

Contour plots of gas volume fraction at 3.75 rps for gas-liquid system:
(a) Axial cross section. (b) Radial cross section at bottom RT (0.15 m). (c) Radial cross section at middle PBTD (0.45 m). (d) Radial cross section at top PBTD (0.75 m).
Figure 4(d, e, f) gives contour plots of the gas-phase velocity magnitude profiles at impeller rotation speed 350 rpm. Compared to Figure 6 in Ford et al. 2008, similar patterns can be observed, and the impeller is loaded. The gas volume fraction is still high near the impeller (see Figure S9(e) in Supplementary Materials). However, a high gas volume fraction region can be found near the impeller tips, which indicates bubbles are beginning to be dispersed. As seen in Figure 6 of (Ford et al. 2008), a ring pattern can be observed in Figure 4, which means that gas is going radially and around the impeller. As is typical of the loaded regime, a well-dispersed gas phase can be expected, while the gas volume fraction is low below the impeller. Similar to air flow pattern, radial-pumping flow patterns is observed in the liquid phase (see Figure S4(d–f) in Supplementary Materials) as well.

Flow patterns of gas phase at 3.75 rps for gas-liquid system: (a) 2D velocity vectors on the axial cross section. (b) Contour pot of magnitude of velocity on the axial cross section. (c) Contour plot of axial velocity on the axial cross section. (d–f) Contour plots of magnitude of velocity on the radial cross section at bottom RT (d), middle PBTD (e), and top PBTD (f). (g–i) contour plots of axial velocity on the radial cross section at bottom RT (g), middle PBTD (h), and top PBTD (i).
Contour plots of the gas-phase velocity magnitude profiles at impeller rotation speed 700 rpm are shown in Figure 4(g, h, i). Similar patterns to Figure 7 in (Ford et al. 2008) can be observed, and the flow is in the completely dispersed regime so that the gas is well dispersed throughout the reactor. (Figure S4(g, h, i) in the Supplementary Materials shows the flow pattern of water at impeller rotation speed 700 rpm.) Like the air velocity profile at this rotation speed, a radial-pumping flow pattern for liquid phase is obtained.

Contour plots of gas volume fraction at 5.08 rps for gas-liquid system:
(a) Axial cross section. (b) Radial cross section at bottom RT (0.15 m). (c) Radial cross section at middle PBTD (0.45 m). (d) Radial cross section at top PBTD (0.75 m).
4.2 Gas-liquid simulations of reactor 2
Table 4 lists the gassed impeller power consumption PG in the tank at two different impeller rotation speeds (3.75 and 5.08 rps) found from the simulations. The comparison of fractional gas holdup, εG, at the corresponding PG both for experiments and simulations are also given in Table 4.
Fractional gas holdup.
Fractional gas holdup | |||
---|---|---|---|
Rotation speed (rps) | log (PG/V) W/m3 | Experimental | Simulation |
3.75 | 1.9356 | 0.0240 | 0.0360 |
5.08 | 1.8062 | 0.0223 | 0.0306 |
For the experimental fractional gas holdup values, in Shewale and Pandit (2006) a power-law type correlation as a function of
where Po and QG are the ungassed impeller power consumption and gas flow rate. Although a similar correlation is proposed in (Michael and Miller 1962) for single disc turbine, α and n parameters are added in Shewale and Pandit (2006) as fitting parameters for PBTD and are given in Table 2 of (Shewale and Pandit 2006).
On the other hand, in the simulations the gassed impeller power consumption in the vessel (PG) is calculated by multiplying the magnitude of the torque (T) by the angular velocity
where NR is the impeller rotational speed. The torque is calculated by integrating the cross product of the total stress
where n is the normal vector of cell face on the surface of shaft and impellers, and total stress
with stress tensors τg and τl being calculated using Eq. (10). It is unclear whether the results from Eq. (17) can be compared quantitatively to the correlation in Eq. (16) which is based on the total power consumption.
Figure 5 gives contour plots of the gas volume fraction in Reactor 2 at rotation speed 3.75 rps. Results indicate that the gas is dispersed effectively at the middle and top PBTDs, but not so well below and above the bottom RT, which agrees with the picture of Effective dispersion-Flooding (DE-F) regime shown in Shewale and Pandit (2006). A ring pattern can be found at the bottom RT, which means gas is going radially and around the impeller, but the impeller cannot disperse gas very effectively to the region away from the impeller. The flow pattern of air at impeller rotation speed 3.75 rps is shown in Figure 6. Air is mainly going upward from the sparger to the top of the tank. At the bottom RT, air is going up around the impeller. Upward and downward circulations are barely formed below and above the impeller. At the middle and top PBTDs, downward circulations can only be observed very close to the impellers.
Radial-pumping and down-pumping flow patterns of water are found at the bottom RT, and air is going up around the impeller (see Figure S10 in Supplementary Materials). Upward and downward circulations are barely formed below and above the impeller. At the middle and top PBTDs, downward circulations can only be observed very close to the impellers. Radial-pumping and down-pumping flow patterns of water at the bottom RT and upper PBTDs at 3.75 rps are more obvious than the flow patterns of air, shown in Figure 5. At the bottom RT, an upward and a downward circulation can be seen clearly below and above the impeller, and water is moving downward near the tank wall. At the middle PBTD, a downward circulation is formed, but still close to impeller region. An upward circulation is also formed between the bottom RT and middle PBTD. At the top PBTD, a downward circulation if formed, and expanded to regions away from the impeller.
Figure 7 shows contour plots of gas volume fraction at rotation speed 5.08 rps. Gas is dispersed effectively at the upper PBTDs, and dispersed better above the bottom RT, compared to gas dispersion at rotation speed 3.75 rps. At the bottom RT, a high gas volume fraction region can also be observed near the tank wall and baffles. Simulation results are similar to the photo shown in Shewale and Pandit (2006), described as the Effective-Loading (DE-L) regime.
The air flow pattern at impeller rotation speed 5.08 rps shown in Figure 8. Like the flow pattern of air at 3.75 rps, air is mainly going upward. A radial-pumping flow pattern can be observed at the bottom RT, although the upward and downward circulations are very subtle below and above the impeller. At the middle and top PBTDs, downward circulations are formed close to the impellers (Figure S11 in the Supplementary Materials shows the water flow pattern at rotation speed of 5.08 rps). A radial-pumping flow pattern is clear at the bottom RT. Below the bottom RT, an upward circulation is formed, while above the impeller, a downward circulation can be observed. At the upper PBTDs, downward circulations that are expanded to regions away from the impellers are generated.

Flow patterns of gas phase at 5.08 rps for gas-liquid system: (a) 2D velocity vectors on the axial cross section. (b) Contour pot of magnitude of velocity on the axial cross section. (c) Contour plot of axial velocity on the axial cross section. (d–f) Contour plots of magnitude of velocity on the radial cross section at bottom RT (d), middle PBTD (e), and top PBTD (f). (g–i) Contour plots of axial velocity on the radial cross section at bottom RT (g), middle PBTD (h), and top PBTD (i).
4.3 Gas-liquid-solid simulations of reactor 2
Contour plots of gas, solid and liquid volume fractions at impeller rotation speed 5.08 rps are shown in Figure 9. The addition of solid particles changes the volume fraction profile significantly. Air is sucked into the mixture near the free surface. At the bottom RT, the ring pattern of gas volume fraction disappears, and the impeller tends to become flooded. A high gas volume fraction is observed near the shaft and impellers, while outside the impeller regions gas is not dispersed effectively. Accordingly, solid and water volume fractions are low near the shaft and impellers. For the solid phase, the particle concentration increases slightly with decreasing distance from the bottom of the vessel.

Contour plots of gas (left), solid (middle) and liquid (right) volume fractions at 5.08 rps: (A–C) axial cross section. (D–F) radial cross section at bottom RT. (G–I) radial cross section at middle PBTD. (J–L) radial cross section at top PBTD. (A, D, G, J) gas phase. (B, E, H, K) solid phase. (C, F, I, L) liquid phase.
With the addition of a solid phase, the radial-pumping flow pattern at the bottom RT seen in the gas-liquid system disappears, and downward circulations at the upper PBTDs can barely be observed. Air mostly goes upward throughout the vessel, and only goes downward in a small region between the blades of the upper PBTDs. Compared to the gas-liquid system, the water flow pattern changes dramatically with the addition of solid particles (See Figure S12 in Supplementary Materials). The radial-pumping flow pattern at the bottom RT cannot be observed anymore. Instead, an upward circulation is formed near the impeller. The upper PBTDs present a radial-pumping more than a down-pumping flow pattern. Upward and downward circulations are formed below and above both impellers. Furthermore, solid particles follow the liquid, and therefore the solid phase has a similar flow pattern to the liquid phase.
5 Conclusions
Multiphase CFD simulations were performed for two STRs. In Reactor-1, gas-liquid simulations were performed. Similar flow patterns with experiments in Ford et al. (2008) are found and the simulation gas holdup values are in good agreement with experimental overall gas holdup values. In Reactor-2, the flow patterns from gas-liquid simulations agree well with experiments (Shewale and Pandit 2006), however, calculated gassed power consumption is smaller than the experimental value. When solid particles added to the Reactor-2 simulations, the bottom RT tends to be flooded, while a radial-pumping flow pattern is observed at the upper PBTDs instead of down-pumping flow.
Acknowledgments
X. Hu, A. Passalacqua and R. O. Fox gratefully acknowledge financial support from SABIC Geleen, The Netherlands.
Author contributions: All the authors have accepted responsibility for the entire content of this submitted manuscript and approved submission.
Research funding: None declared.
Conflict of interest statement: The authors declare no conflicts of interest regarding this article.
References
Achouri, R., I. Mokni, H. Mhiri, and P. Bournot. 2012. “A 3D CFD Simulation of a Self-Inducing Pitched Blade Turbine Downflow.” Energy Conversion and Management 64: 633–41. https://doi.org/10.1016/j.enconman.2012.06.005.Suche in Google Scholar
Ahmed, S. U., P. Ranganathan, A. Pandey, and S. Sivaraman. 2010. “Computational Fluid Dynamics Modeling of Gas Dispersion in Multi Impeller Bioreactor.” Journal of Bioscience and Bioengineering 109: 588–97. https://doi.org/10.1016/j.jbiosc.2009.11.014.Suche in Google Scholar
Arlov, D., J. Revstedt, and L. Fuchs. 2008. “Numerical Simulation of a Gas-Liquid Rushton Stirred Reactor LES and LPT.” Computers & Fluids 37: 793–801. https://doi.org/10.1016/j.compfluid.2007.03.017.Suche in Google Scholar
Bakker, A., and H. E. A. Van Den Akker. 1994. “A Computational Model for the Gas-Liquid Flow in Stirred Reactors.” Chemical Engineering Research & Design 72: 594–606.Suche in Google Scholar
Boyer, C., A. M. Duquennt, and G. Wild. 2002. “Measuring Techniques in Gas-Liquid and Gas-Liquid-Solid Reactors.” Chemical Engineering Science 57: 31853215. https://doi.org/10.1016/s0009-2509(02)00193-8.Suche in Google Scholar
Burns, A. D., T. Frank, I. Hamill, and J. M. Shi. 2004. “The Favre Averaged Drag Model for Turbulent Dispersion in Eulerian Multi-Phase Flows.” In 5th International Conference on Multiphase Flow. ICMF’04 Yokohama, Japan, May 30–June 4, Paper No. 392. Yokohama, Japan: ICMFSuche in Google Scholar
Chen, L., Y. Bao, and Z. Gao. 2009. “Void Fraction Distribution in Cold-Gassed and Hot-Sparged Three Phase Stirred Tanks with Multi-Impeller.” Chinese Journal of Chemical Engineering 17: 887–95. https://doi.org/10.1016/s1004-9541(08)60293-1.Suche in Google Scholar
Conway, K., A. Kyle, and C. D. Rielly. 2002. “Gas-Liquid-Solid Operation of a Vortex Ingesting Stirred Tank Reactor.” Chemical Engineering Research and Design 80: 839–45. https://doi.org/10.1205/026387602321143372.Suche in Google Scholar
Dohi, N., Y. Matsuda, K. Shimizu, K. Minekawa, and Y. Kawase. 2002. “An Experimental Investigation into Vapor Dispersion and Solid Suspension in Boiling Stirred Tank Reactors.” Chemical Engineering and Processing: Process Intensification 41: 267–79. https://doi.org/10.1016/s0255-2701(01)00142-8.Suche in Google Scholar
Dohi, N., T. Takahashi, K. Minekawa, and Y. Kawase. 2004. “Power Consumption and Solid Suspension Performance of Large-Scale Impellers in Gas-Liquid-Solid Three-Phase Stirred Tank Reactors.” Chemical Engineering Journal 97: 103–14. https://doi.org/10.1016/s1385-8947(03)00148-7.Suche in Google Scholar
Drew, D. A. 1983. “Mathematical Modeling of Two-Phase Flow.” Annual Review of Fluid Mechanics 15: 261–91. https://doi.org/10.1146/annurev.fl.15.010183.001401.Suche in Google Scholar
Drew, D. A., and R. T. Lahey. 1987. “The Virtual Mass and Lift Force on a Sphere in Rotating and Straining Inviscid Flow.” International Journal of Multiphase Flow 13: 113–21. https://doi.org/10.1016/0301-9322(87)90011-5.Suche in Google Scholar
Dutta, N. N., and V. G. Pangarkar. 1995. “Critical Impeller Speed for Solid Suspension Un Multi-Impeller Three Phase Agitated Contactors.” The Canadian Journal of Chemical Engineering 73: 273–83. https://doi.org/10.1002/cjce.5450730302.Suche in Google Scholar
Fishwick, R. P., J. M. Winterbottom, and E. H. Stitt. 2003. “Effect of Gassing Rate on Solid-Liquid Mass Transfer Coefficients and Particle Slip Velocities in Stirred Tank Reactors.” Chemical Engineering Science 58: 1087–93. https://doi.org/10.1016/s0009-2509(02)00651-6.Suche in Google Scholar
Ford, J. J., T. J. Heindel, T. C. Jensen, and J. B. Drake. 2008. “X-Ray Computed Tomography of a Gas-Sparged Stirred-Tank Reactor.” Chemical Engineering Science 63: 2075–85. https://doi.org/10.1016/j.ces.2008.01.007.Suche in Google Scholar
Gosman, A. D., C. Lekakou, S. Politis, R. I. Issa, and M. K. Looney. 1992. “Multidimensional Modeling of Turbulent Two-Phase Flows in Stirred Vessels.” AIChE Journal 38: 1946–56. https://doi.org/10.1002/aic.690381210.Suche in Google Scholar
Himmelsbach, W., D. Houlton, D. Ortlieb, and M. Lovallo. 2006. “New Advances in Agitation Technology for Exothermic Reactions in Very Large Reactors.” Chemical Engineering Science 61: 3044–52. https://doi.org/10.1016/j.ces.2005.10.059.Suche in Google Scholar
Jahoda, M., M. Mostek, A. Kukukukova, and V. Machon. 2007. “CFD Modelling of Liquid Homogenization in Stirred Tanks with One and Two Impellers Using Large Eddy Simulation.” Chemical Engineering Research & Design 85: 616–25. https://doi.org/10.1205/cherd06183.Suche in Google Scholar
Jahoda, M., L. Tomaskove, and M. Moste. 2009. “CFD Prediction of Liquid Homogenisation in a Gas-Liquid Stirred Tank.” Chemical Engineering Research and Design 87: 460–7. https://doi.org/10.1016/j.cherd.2008.12.006.Suche in Google Scholar
Joshi, J. B., N. K. Nere, C. V. Rane, B. N. Murthy, and C. S. Mathpati. 2011a. “CFD Simulation of Stirred Tanks: Comparison of Turbulence Models (Part 1: Radial Flow Impellers).” The Canadian Journal of Chemical Engineering 89: 23–82. https://doi.org/10.1002/cjce.20446.Suche in Google Scholar
Joshi, J. B., N. K. Nere, C. V. Rane, B. N. Murthy, C. S. Patwardhan, A. W. Mathpati, and V. V. Ranade. 2011b. “CFD Simulation of Stirred Tanks: Comparison of Turbulence Models (Part II: Axial Flow Impellers, Multiple Impellers and Multiphase Dispersions).” The Canadian Journal of Chemical Engineering 89: 754–816. https://doi.org/10.1002/cjce.20465.Suche in Google Scholar
K’alal, Z., M. Jahoda, and I. Fort. 2014. “CFD Prediction of Gas-Liquid Flow in an Aerated Stirred Vessel Using the Population Balance Model.” Chemical and Process Engineering 35: 55–73. https://doi.org/10.2478/cpe-2014-0005.Suche in Google Scholar
Kasat, G. R., and A. B. Pandit. 2004. “Mixing Time Studies in Multiple Impeller Agitated Reactors.” The Canadian Journal of Chemical Engineering 82: 892–904. https://doi.org/10.1002/cjce.5450820504.Suche in Google Scholar
Kerdouss, F., A. Bannaari, and P. Proulx. 2006. “CFD Modeling of Gas Dispersion and Bubble Size in a Double Turbine Stirred Tank.” Chemical Engineering Science 61: 3313–22. https://doi.org/10.1016/j.ces.2005.11.061.Suche in Google Scholar
Khopkar, A. R., and P. A. Tanguy. 2008. “CFD Simulation of Gas-Liquid Flows in Stirred Vessel Equipped with Dual Rushton Turbines: Influence of Parallel, Merging and Diverging Flow Configurations.” Chemical Engineering Science 63: 3810–20. https://doi.org/10.1016/j.ces.2008.04.039.Suche in Google Scholar
Khopkar, A. R., G. R. Kasat, A. B. Pandit, and V. V. Ranade. 2006. “CFD Simulation of Mixing in Tall Gas-Liquid Stirred Vessel: Role of Local Flow Patterns.” Chemical Engineering Science 61: 2921–9. https://doi.org/10.1016/j.ces.2005.09.023.Suche in Google Scholar
Kumaresan, T., and J. B. Joshi. 2006. “Effect of Impeller Design on the Flow Pattern and Mixing in Stirred Tanks.” Chemical Engineering Journal 115: 173–93. https://doi.org/10.1016/j.cej.2005.10.002.Suche in Google Scholar
Lane, G. L., M. P. Schwarz, and G. M. Evans. 2005. “Numerical Modelling of Gas-Liquid Flow in Stirred Tanks.” Chemical Engineering Science 60: 2203–14. https://doi.org/10.1016/j.ces.2004.11.046.Suche in Google Scholar
Luo, J. Y., Y. Issa, and A. D. Gosman. 1994. “Prediction of Impeller-Induced Flow in Mixing Vessels Using Multiple Frames of References.” Institution of Chemical Engineering Symposium Series 136: 549–56.Suche in Google Scholar
Micale, G., V. Carrara, F. Grisafi, and A. Brucato. 2000. “Solids Suspension in Three-Phase Stirred Tanks.” Chemical Engineering Research and Design 78: 319–26. https://doi.org/10.1205/026387600527374.Suche in Google Scholar
Michael, B. J., and S. A. Miller. 1962. “Power Requirements of Gas-Liquid Agitated Systems.” AIChE Journal 8: 262–6. https://doi.org/10.1002/aic.690080226.Suche in Google Scholar
Miller, D. N. 1974. “Scale-Up of Agitated Vessels Gas-Liquid Mass Transfer.” AIChE Journal 20: 445–53. https://doi.org/10.1002/aic.690200303.Suche in Google Scholar
Min, J., Y. Bao, L. Chen, Z. Gao, and J. M. Smith. 2008. “Numerical Simulation of Gas Dispersion in an Aerated Stirred Reactor with Multiple Impellers.” Industrial & Engineering Chemistry Research 47: 7112–7. https://doi.org/10.1021/ie800490j.Suche in Google Scholar
Moilanen, P., M. Laakkonen, O. Visuri, V. Alopaeus, and J. Aittamaa. 2008. “Modelling Mass Transfer in an Aerated 0.2 M3 Vessel Agitated by Rushton, Phasejet and Combijet Impellers.” Chemical Engineering Journal 142: 95–108. https://doi.org/10.1016/j.cej.2008.01.033.Suche in Google Scholar
Murthy, B. N., and J. B. Joshi. 2008. “Assessment of Standard K-Epsilon, RSM and LES Turbulence Models in a Baffled Stirred Vessel Agitated by Various Impeller Designs.” Chemical Engineering Science 63: 5468–95. https://doi.org/10.1016/j.ces.2008.06.019.Suche in Google Scholar
Murthy, B. N., R. S. Ghadge, and J. B. Joshi. 2007. “CFD Simulations of Gas-Liquid-Solid Stirred Reactor: Prediction of Critical Impeller Speed for Solid Suspension.” Chemical Engineering Science 62: 7184–95. https://doi.org/10.1016/j.ces.2007.07.005.Suche in Google Scholar
Nienow, A. W. 1998. “Hydrodynamics of Stirred Bioreactors.” Applied Mechanics Reviews 51: 3–32. https://doi.org/10.1115/1.3098990.Suche in Google Scholar
Nienow, A. W., and W. Bujalski. 2002. “Recent Studies on Agitated Three-Phase (Gas-Solid-Liquid) Systems in the Turbulent Regime.” Chemical Engineering Research and Design 80: 832–8. https://doi.org/10.1205/026387602321143363.Suche in Google Scholar
Nienow, A. W., M. C. G. Warmoeskerken, J. M. Smith, and M. Konno. 1985. “On the Flooding/Loading Transition and the Completer Dispersal Condition in Aerated Vessels Agitated by a Rushton Turbine.” In Proceedings of the 5th European Conference on Mixing, 143–54. Wurzburg: BHRA Cranfield.Suche in Google Scholar
OpenFOAM. 2015. The Open Source CFD Toolbox – User’s Guide. United Kingdom: OpenCFD Ltd.Suche in Google Scholar
Paglianti, A. 2002. “Simple Model to Evaluate Loading/Flooding Transition in Aerated Vessels Stirred by Rushton Disc Turbines.” The Canadian Journal of Chemical Engineering 80: 1–5. https://doi.org/10.1002/cjce.5450800409.Suche in Google Scholar
Panneerselvam, R., S. Savithri, and G. D. Surender. 2009. “Computational Fluid Dynamics Simulation of Solid Suspension in a Gas-Liquid-Solid Mechanically Agitated Contactor.” Industrial & Engineering Chemistry Research 48: 1608–20. https://doi.org/10.1021/ie800978w.Suche in Google Scholar
Parasu-Veera, U., A. W. Patwardhan, and J. B. Joshi. 2001. “Measurement of Gas Hold-Up Profiles in Stirred Tank Reactors by Gamma Ray Attenuation Technique.” Chemical Engineering Research and Design 79: 684–8. https://doi.org/10.1205/026387601316971352.Suche in Google Scholar
Petitti, M., M. Vanni, D. L. Marchisio, A. Buffo, and F. Podenzani. 2013. “Simulation of Coalescence, Break-Up and Mass Transfer in a Gas-Liquid Stirred Tank with CQMOM.” Chemical Engineering Journal 228: 1182–94. https://doi.org/10.1016/j.cej.2013.05.047.Suche in Google Scholar
Ranade, V. V., and H. E. A. Van Den Akker. 1994. “A Computational Snapshot of Gas-Liquid Flow in Baffled Stirred Reactors.” Chemical Engineering Science 49: 5175–92. https://doi.org/10.1016/0009-2509(94)00318-1.Suche in Google Scholar
Rewatkar, V. B., and J. B. Joshi. 1992. “Effect of Addition of Alcohol on the Design Parameters of Mechanically Agitated Three-Phase Reactors.” The Chemical Engineering Journal 49: 107–17. https://doi.org/10.1016/0300-9467(92)80044-b.Suche in Google Scholar
Rewatkar, V. B., A. J. Deshpande, A. B. Pandit, and J. B. Joshi. 1993. “Gas Hold-Up Behavior of Mechanically Agitated Gas-Liquid Reactors Using Pitched Blade Downflow Turbines.” The Canadian Journal of Chemical Engineering 71: 226–37. https://doi.org/10.1002/cjce.5450710209.Suche in Google Scholar
Sawant, S. B., and J. B. Joshi. 1979. “Critical Impeller Speed for the Onset of Gas Induction in Gas-Inducing Types of Agitated Contactors.” The Chemical Engineering Journal 18: 87–91. https://doi.org/10.1016/0300-9467(79)80018-0.Suche in Google Scholar
Scargiali, F., A. D’Orazio, F. Grisafi, and A. Brucato. 2007. “Modelling and Simulation of Gas-Liquid Hydrodynamics in Mechanically Stirred Tanks.” Chemical Engineering Research & Design 85: 637–46. https://doi.org/10.1205/cherd06243.Suche in Google Scholar
Shewale, S. D., and A. B. Pandit. 2006. “Studies in Multiple Impeller Agitated Gas-Liquid Contractors.” Chemical Engineering Science 61: 486–504. https://doi.org/10.1016/j.ces.2005.04.078.Suche in Google Scholar
Smagorinsky, J. 1963. “General Circulation Experiments with the Primitive Equations.” Monthly Weather Review 91: 99–164. https://doi.org/10.1175/1520-0493(1963)091<0099:gcewtp>2.3.co;2.10.1175/1520-0493(1963)091<0099:GCEWTP>2.3.CO;2Suche in Google Scholar
Tomiyama, A., I. Kataoka, I. Zun, and T. Sakaguchi. 1998. “Drag Coefficients of Single Bubbles under Normal and Micro Gravity Conditions.” JSME International Journal Series B 41: 472–9. https://doi.org/10.1299/jsmeb.41.472.Suche in Google Scholar
Tyagi, M., S. Roy, A. D. HarveyIII, and S. Acharya. 2007. “Simulation of Laminar and Turbulent Impeller Stirred Tanks Using Immersed Boundary Method and Large Eddy Simulation Technique in Multi-Block Curvilinear Geometries.” Chemical Engineering Science 62: 1351–63. https://doi.org/10.1016/j.ces.2006.11.017.Suche in Google Scholar
Vr’abel, P., R. G. J. M. Van der Lans, Y. Q. Cui, and K. C. A. M. Luyben. 1999. “Compartment Model Approach: Mixing in Large Scale Aeraated Reactors with Multiple Impellers.” Chemical Engineering Research and Design 77: 291–302. https://doi.org/10.1205/026387699526223.Suche in Google Scholar
Vr’abel, P., R. G. J. M. Van der Lans, K. C. A. M. Luyben, L. Boon, and A. W. Nienow. 2000. “Mixing in Large-Scale Vessels Stirred with Multiple Radial or Radial and Axial Up-Pumping Impellers: Modelling and Measurements.” Chemical Engineering Science 55: 5881–96. https://doi.org/10.1016/S0009-2509(00)00175-5.Suche in Google Scholar
Wen, C. Y., and Y. H. Hu. 1966. “Mechanics of Fluidization.” Chemical Engineering Progress Symposium Series 62: 100–11.Suche in Google Scholar
You, S. T., A. A. A. Raman, R. S. S. R. E. Shah, and N. M. I. Mohamad. 2014. “Multiple-Impeller Stirred Vessel Stuides.” Reviews in Chemical Engineering 30: 323–36. https://doi.org/10.1515/revce-2013-0028.Suche in Google Scholar
Supplementary Material
The online version of this article offers supplementary material (https://doi.org/10.1515/ijcre-2019-0229).
© 2021 Xiaofei Hu et al., published by De Gruyter, Berlin/Boston
This work is licensed under the Creative Commons Attribution 4.0 International License.
Artikel in diesem Heft
- Frontmatter
- Articles
- Mixing performance of an electroosmotic micromixer with Koch fractal structure
- Minimum time controller in a class of chemical reactors based on Lagrangian approach
- Better efficiency for the olefin plant demethanizer tower by replacing trays with packing
- CFD-PBM simulation of hydrodynamics of microbubble column with shear-thinning fluid
- Predictive model development and simulation of photobioreactors for algal biomass growth estimation
- Kinetic analysis of dual impellers on methane hydrate formation
- CFD simulation of ultrasonic atomization pyrolysis reactor: the influence of droplet behaviors and solvent evaporation
- Mesoporous KIT-6 supported Cr and Co-based catalysts for microwave-assisted non-oxidative ethane dehydrogenation
- CFD simulations of stirred-tank reactors for gas-liquid and gas-liquid-solid systems using OpenFOAM®
Artikel in diesem Heft
- Frontmatter
- Articles
- Mixing performance of an electroosmotic micromixer with Koch fractal structure
- Minimum time controller in a class of chemical reactors based on Lagrangian approach
- Better efficiency for the olefin plant demethanizer tower by replacing trays with packing
- CFD-PBM simulation of hydrodynamics of microbubble column with shear-thinning fluid
- Predictive model development and simulation of photobioreactors for algal biomass growth estimation
- Kinetic analysis of dual impellers on methane hydrate formation
- CFD simulation of ultrasonic atomization pyrolysis reactor: the influence of droplet behaviors and solvent evaporation
- Mesoporous KIT-6 supported Cr and Co-based catalysts for microwave-assisted non-oxidative ethane dehydrogenation
- CFD simulations of stirred-tank reactors for gas-liquid and gas-liquid-solid systems using OpenFOAM®