You are on page 1of 12

Chemical Engineering Science 92 (2013) 112

Contents lists available at SciVerse ScienceDirect

Chemical Engineering Science


journal homepage: www.elsevier.com/locate/ces

Recent developments in experimental (PIV) and numerical (DNS) investigation of solidliquid uidized beds
R.K. Reddy a,c, M.J. Sathe a,d, J.B. Joshi a,b,n, K. Nandakumar c,nn, G.M. Evans d
a

Department of Chemical Engineering, Institute of Chemical Technology, Matunga, Mumbai 400 019, India Homi Bhabhi National Institute, Anushakti Nagar, Mumbai 400 094, India c Department of Chemical Engineering, Louisiana State University, Baton Rouge, LA 70803, USA d School of Engineering, The University of Newcastle, Callaghan, NSW 2308, Australia
b

H I G H L I G H T S
c c c c c

Particle image velocimetry and DNS (ctitious-domain method) have been performed. The particle Reynolds number range was varied from 51 to 759. Refractive index of uid is matched with particles to achieve transparency for PIV. Space and time mean and RMS velocity eld were observed to be homogeneous. Energy dissipation rate close to particles exceeds the average in uidized bed.

a r t i c l e i n f o
Article history: Received 4 February 2012 Received in revised form 3 November 2012 Accepted 14 November 2012 Available online 20 November 2012 Keywords: CFD Direct numerical simulation Particle image velocimetry Fluidized bed Hindrance effect Multiphase ow

abstract
Particle image velocimetry measurements have been performed in a solidliquid uidized bed in the Reynolds number range 51759. To do this, the refractive indexes of the solid and liquid phases were matched at approximately 1.47 using 3 mm diameter borosilicate glass beads and a solution of turpentine and tetra-hydronaphthalene. Parafn oil was added in varying quantities to vary the dynamic viscosity between 0.0012 and 0.010 Pa s without changing the refractive index of the solution. From the PIV measurements, at sampling rates of 2 Hz, the uctuating velocity components were found to be quite uniform in both the axial and radial directions. Moreover, the computed turbulent kinetic energy dissipation rates were also found to be relatively constant throughout the bed, thus highlighting the homogenous nature of the turbulence within the system. Following from Reddy et al. (2010b), direct numerical simulations were undertaken at particle Reynolds numbers up to 200 for assemblages of 1, 9, 27, 100, 180 and 245 particles, which corresponded to a liquid volume fraction range of 0.687 o AL o 0.998. The effect of surrounding particles on the settling velocity (hindrance effect) and the wake dynamics was investigated. It was found that the average settling velocity decreased with an increasing number of particles, with the quantitative results being in good agreement with the well established empirical correlation of Richardson and Zaki (1954). The local energy dissipation rate was also computed, and for a particle Reynolds number of 51, it was found to be 5.5 m2 s 3. This value was approximately 18 times the average energy dissipation rate of 0.30 m2 s 3; and compared favourably with the 0.36 m2 s 3obtained by a volume-averaged energy balance of the experimental system. & 2012 Elsevier Ltd. All rights reserved.

1. Introduction Particleuid interaction plays an important role in many industrial processes, including: crystallization, chromatographic separations, ion exchange, sedimentation, uidization, hydrometallurgical operations (leaching, particle classication and backwashing of downow granular lters), slurry transport (water lubricated transport of heavy crude and coal slurries), hydraulic fracturing in oil and natural gas production. Whilst

n Corresponding author at: Department of Chemical Engineering, Institute of Chemical Technology, Matunga, Mumbai 400 019, India. Tel.: 91 22 3361 2106; fax: 91 22 3361 1020. nn Corresponding author. Tel.: 1 225 578 2361; fax: 1 225 578 1476. E-mail addresses: jbjoshi@gmail.com (J.B. Joshi), nandakumar@lsu.edu (K. Nandakumar).

0009-2509/$ - see front matter & 2012 Elsevier Ltd. All rights reserved. http://dx.doi.org/10.1016/j.ces.2012.11.017

R.K. Reddy et al. / Chemical Engineering Science 92 (2013) 112

these multiphase processes are widely applied detailed understanding of the relative motion between the particles and the uid and their corresponding inuence on mass, heat and momentum transfer [for instance, Joshi et al., 1980, 1981, 2003; Thakre and Joshi, 1999; Murthy et al., 2007] has been limited for two reasons. First, non-intrusive experimental observations are very difcult to undertake, especially of the liquid motion in the interstices between the particles, droplet or bubbles. Second, the computational requirements for direct numerical modelling of such a complex ow have been prohibitively large. Recent advances in both ow visualization techniques and computing capability have reduced these limitations to such an extent that detailed investigation of multiphase systems that more closely reect actual industrial practices is now possible. In this study, particle image velocimetry (PIV) has been utilized to quantify both the particle slip and interstitial liquid velocity within a solidliquid uidized bed in the Reynolds number range of 51759. The experimental measurements are complemented by in-house direct numerical simulations (DNS) to quantify the inuence of the presence of neighbouring particles on the wake structure and settling velocity of freely falling particles. The DNS is carried out using the in-house code reported in Jin et al. (2009), whereby the hydrodynamic force between the particle and the uid is resolved without the need for assuming a given drag law relationship. 1.1. Previous work Numerous experimental and theoretical studies (e.g. Richardson and Zaki, 1954; Hanratty and Bandukwala, 1957; Garside and A1-Dibouni, 1977; Joshi, 1983; Pandit and Joshi, 1998; Gevrin et al., 2008; Reddy and Joshi., 2009; Zivkovic et al., 2009) have investigated bed expansion characteristics of solidliquid uidized beds; and whilst these studies are useful they give no real insight into the ow behaviour in the interstices between the particles. There have been a number of studies that have attempted to address this issue. Handely et al. (1966) undertook photographic measurements in beds of 3 mm glass beads at ReN of 45, 87 and 182 in a 75 mm diameter glass column and found that the root mean square uctuating velocity in the axial direction is about 2.5 times higher than that in the radial direction. Chen and Kadambi (1990) studied solidliquid slurry ows in a horizontal pipe using Laser Doppler Velocimetry (LDV) and refractive index matching of the solid and liquid phases. They used silica-gel particles having an average size of 40 mm and 50% W/W sodium iodide aqueous solution. They were able to measure the average axial liquid velocity for silica-gel concentrations between 5 and 50% W/W. Neither Handely et al. (1966) nor Chen and Kadambi (1990) reported turbulence quantities. Chen and Fan (1992) applied PIV to three phase gasliquid solid uidized beds to obtain 49 velocity vectors that provided an excellent basis for quantifying the mean ow. However, the number was well below the $100,000 required for reliable estimation of turbulence intensity, turbulent kinetic energy, turbulent energy dissipation rate and structure functions. Haam et al. (2000) combined PIV measurements with refractive index matching to obtain velocity information inside solidliquid uidized beds at solids concentrations much higher than was previously possible. They measured the axial and radial velocity components in a uidized bed of glass beads for ReN values of 1040 and 1550. They found that the axial turbulent intensity of the uid increased by up to 70% due to the presence of the glass beads. Dijkhuizen et al. (2007) coupled their PIV measurements with simultaneous measurement of the instantaneous velocity and granular temperature elds. The PIV algorithm was specically

optimized for dense granular systems and applied to a uidized bed at incipient uidization conditions into which both singleand multi-bubbles were injected. They observed that the highest granular temperature was in the vicinity of the bubble(s), and for the case of 1.5 mm glass particle bed the granular temperature (Gidaspow, 1994) was in the range of 0.0220.069 m2 s 2. Shi (2007) applied PIV to investigate the particle motion and cluster properties in a gassolid two-phase ow in a circulating uidized bed riser. Visual images and micro-structure of various clusters were captured. After the boundary of clusters was determined by the gray level threshold method, clusters were classied by the distance between particles and the shape and position of clusters. In addition, the process of cluster formation and breakup was described, and the sizes of clusters were also obtained. With the Minimum Quadric Difference cross-correlation algorithm suitable for high-density particles, the axial velocities of the particles were obtained in the dilute phase section. Analysis of the magnitude and distribution of particle axial velocity in the radial direction showed at most radial cross-sections a parabolic prole in the upward direction. The magnitude of axial velocity in the core region was found to be higher than that in the near wall region of the riser. Muller et al. (2009) employed simultaneous PIV and Planar Laser-Induced Fluorescence (PLIF) measuring techniques to investigate the eruption of both a single and continuous stream of bubbles in the freeboard region of a uidized bed. The observed bubble eruption patterns were in general agreement with the bubble models published in the literature. Based on the calculated vorticity of the gas in the freeboard it was found that the bubble induced turbulence decays rapidly. Stereoscopic PIV measurements of the out-of-plane component of the liquid velocity were found to be not negligible. Kashyap et al. (2011) used PIV to obtain laminar and turbulent properties near the wall in the developing region of circulation of Geldart B type particles in the riser part of circulating gassolid uidized bed. Instantaneous velocities for the solid phase were measured simultaneously in the axial and radial directions using a CCD camera and a coloured rotating transparency. A novel method was used to determine axial and radial solid phase dispersion coefcients using the autocorrelation technique. The measured laminar and Reynolds stresses, laminar and turbulent granular temperatures, laminar and turbulent dispersion coefcients and energy spectra all exhibited anisotropy. The total granular temperatures were in reasonable agreement with the literature values. However, the axial and radial solid dispersion coefcients measured near the wall were slightly lower than the radially averaged values in the literature. Hernandez-Jimenez et al. (2011) investigated both experimentally and computationally the hydrodynamics of a rectangular, bubbling air-uidized bed of 5 mm thickness. The authors used PIV and Digital Image Analysis (DIA) to obtain quantitative information on bubble hydrodynamics, dense-phase probability, and time-averaged vertical and horizontal components of the dense-phase velocity as a function of gas ow rate through the bed. Simulations were also conducted using an EulerEuler twouid approach based on two different closure models for the gas particle interaction, namely the drag models of (1) Gidaspow (1994) and (2) Syamlal and OBrien (1989). Good agreement between experimental observation and simulation results was obtained for dense-phase probability, gas bubble diameter rise velocity. The vertical component of the simulated dense-phase velocity, however, was found to be nearly an order of magnitude larger than that obtained from the PIV experiments. Van Buijtenen et al. (2011) applied PIV, Positron Emission Particle Tracking (PEPT), and Electrical Capacitance Tomography (ECT) measuring techniques to quantify the ow in both quasi-2D

R.K. Reddy et al. / Chemical Engineering Science 92 (2013) 112

and 3D spouted uidized beds. In the pseudo-2D bed the measurements were able to highlight the appearance of dead zones in the annulus near the bottom of the bed in the spoutuidization regime. Measurements were also made in the jet-inuidized-bed regime beds with spout heights of 0, 20, and 40 mm. The results from both experimental systems were compared with full 3D discrete particle modelling (DPM) simulations; with generally good agreement being achieved except for slight overprediction of the velocities in the annular region for the quasi-2D case where wall effects in the experimental system are possibly higher than assumed in the modelling. In addition to the experimental investigations there have been many studies involving simulations of multi-particle systems. For a recent review of the topic, with particular focus on DNS, see Reddy et al. (2010b). Some more recent studies include: Deen et al. (2009) have reported DNS results for multi-uid ows in which simultaneously moving deformable (drops or bubbles) and non-deformable (particles) elements are present, possibly with the additional presence of free surfaces. They utilized a front tracking (FT) model to circumvent the explicit computation of the interface curvature. They also used an immersed boundary (IB) model to incorporate both particleuid and particleparticle interaction via a direct forcing method and a hard sphere Discrete Particle approach. The capabilities of the hybrid FT-IB model are demonstrated by the authors with a number of examples in which complex topological changes in the interface are encountered. Uhlmann and Pinelli (2009) conducted a DNS study of dilute turbulent particulate ow in a vertical plane channel, considering up to 8192 nite-size rigid particles with numerically resolved phase interfaces. The particle diameter corresponded to approximately nine wall units, with a terminal Reynolds number of 136. The bulk uid upow was maintained at a Re 2700 to maintain solids suspension. Two different density ratios were simulated, which varied by a factor of 4.5. The corresponding Stokes numbers for the two particles were O(10) in the near-wall region and O(1) in the outer ow. The DNS simulations indicated that the mean ow velocity prole tended towards a concave shape, and anisotropy for both the turbulence intensity and normal stresses. Large-scale elongated streak-like structures were predicted by the DNS, with stream-wise dimensions of the order of eight channel half-widths and cross-stream dimensions of the order of one channel half-width. There was no evidence for the formation of particle clusters, which suggested that the large structures were due to an intrinsic instability of the ow that was triggered by the presence of the particles. Xu and Subramaniam (2010) have reported DNS results for turbulent ow past uniform and clustered congurations of xed particle assemblies at the same solid volume fraction. Their approach was based on a discrete-time, direct-forcing immersed boundary method that imposes no-slip and no-penetration boundary conditions on each particle surface. Results are reported for mean ow Reynolds number of 50 where the ratio of the particle diameter to the Kolmogorov scale was 5.5. The DNS conrmed experimental observations that the uid-phase turbulent kinetic energy was enhanced by clustered congurations leading to increased levels of anisotropic turbulence. Tenneti et al. (2011) have reported particle-resolved DNS results of interphase momentum transfer in ow past xed random assemblies of mono-disperse spheres with nite uid inertia using a continuum NavierStokes solver. This solver is based on a new formulation called the Particle-resolved Uncontaminated-uid Reconcilable Immersed Boundary Method (PUReIBM). The principal advantage of this formulation is that the uid stress at the particle surface is calculated directly from the ow solution (velocity and pressure elds), which when integrated over the surfaces of all particles yields the average uidparticle force.

The PUReIBM is a consistent numerical method to study gassolid ow because it results in a force density on particle surfaces which is reconcilable with the averaged two-uid theory. The numerical convergence and the accuracy of PUReIBM approach were established through a comprehensive suite of validation tests. The normalized average uidparticle force, F, was obtained as a function of solid volume and mean ow Reynolds number for random assemblies of mono-disperse spheres. From the simulations, a simple correlation for F, and hence drag coefcient, was developed for the particle Re range 100300. Given that the simulations were based on a xed particle assembly, any effects of mobility of the particles was not included. However, approach is a good approximation for high Stokes number particles, which are encountered in most gassolid ows. 1.2. Research priorities The great majority of industrial systems are operated under turbulent conditions, wherein a compendium of eddies (turbulent structures) of different length and time scales govern the momentum, heat and mass transfer and mixing behaviour. For this reason it is desirable to more fully understand the formation and dynamics of these turbulent structures. All the above-mentioned studies on PIV measurement and DNS computations have provided an excellent foundation, and with recent advances in experimental and mathematical techniques there is an opportunity to expand that knowledge for multi-particle systems. The following list of 13 research priorities is aimed at coupling both experimental and computational investigations. The priorities include: 1. DNS simulation for creeping ow (ReN o1) around a single particle in both an innite liquid and a solidliquid uidized bed. Estimation of skin, form and total friction at all locations on the surface of the particle. Estimation of total drag coefcient at different solid volume fractions in the uidized bed to quantify the hindrance effect. 2. DNS simulation of ow around a particle in the ReN range of 1103 to obtain 3D component instantaneous velocities at temporal resolutions up to the Kolmogorov scale. Simulation of boundary layer separation and also the size, shape and stability of wakes behind a particle with respect to ReN. Estimation of skin and form drag on the entire surface of the particle. 3. DNS simulation of ow around a single particle in the ReN range of 103106, capturing the sudden drop in drag coefcient, CD, at ReN of around 2 105 due to the transition from laminar to turbulent boundary layer separation. 4. DNS simulation of settling of a particle in a Hele-Shaw cell to include the wall effect. 5. DNS simulation of ow around two neighbouring particles. 6. DNS simulation of a solidliquid uidized bed in the ReN range of 1106 to obtain 3D component instantaneous velocities at temporal resolutions up to a small scale permissible by the computational power. Estimation of skin, form and total friction at all locations on particles within the uidized bed. Simulation of boundary layer separation in multiparticle systems and also understanding of size, shape and stability of wakes with respect to ReN and solids volume fraction. 7. DNS simulation of solidliquid uidized bed where the momentum transfer is accompanied by mass/heat transfer at the wall. Quantication of heat and mass transfer coefcients as a function of ReN and ALutilizing u0 u0 , u0 c0 and u0 T0 simulated results. 8. Quantication of homogeneity, isotropy and energy spectrum in solidliquid dispersions as a function of ReN, AL and particle diameter and shape.

R.K. Reddy et al. / Chemical Engineering Science 92 (2013) 112

9. Development of scaling laws for the inertial, dissipation and large scale sub-range in solidliquid dispersions. 10. DNS simulation for estimation of interface forces, such as drag, lift, virtual mass, and Basset; and subsequent application in the closure formulation for RANS and LES simulations of solidliquid dispersions. 11. Measurement of instantaneous velocity within the uidized bed using techniques such as high speed PIV, at sampling rates of up to 10 kHz, in combination with refractive index matching of the liquid and solid phases. Application of mathematical techniques, such as discrete wavelet transforms (DWT), continuous wavelet transforms (CWT) and proper orthogonal decomposition (POD), to identify and characterize the ow structures in terms of shape, size, velocity and energy distributions. 12. To understand the mechanism of vorticity generation and hence the origin of turbulence. 13. Relate momentum, heat and mass transfer to turbulent structure dynamics. Priorities 1 and 2 for the authors have already been reported in Reddy et al. (2010a, 2010b). This study is focused on addressing priorities 3, 6 and 11. In terms of DNS (Priorities 3 and 6), simulating up to 245 particles (as performed previously) with Reynolds number range extended to 200. In terms of experimental development (Priority 11), standardization of PIV measurement in terms of matching of refractive indices of the solid particles and liquid; achieving sampling frequencies of 2 Hz over a wide range of Reynolds numbers; and data processing in terms of mean and turbulence characteristics which can be directly compared to DNS results.

Outlet

Gasket

50

Heat 8 Exchanger Acrylic column 1 Outer square column 2

300

Distributor
100

Calming section 4 Flange 5

Vent

Inlet

Refractive index matching liquid


7

PUMP 6
Fig. 1. Solidliquid uidized bed experimental setup (all dimensions are in mm) [(1) acrylic column; (2) outer square column; (3) distributor; (4) calming section; (5) ange; (6) pump; (7) refractive index matching liquid; (8) heat Exchanger; (9) gasket].

2. Experimental 2.1. Solidliquid uidized bed apparatus The schematic of experimental apparatus is shown in Fig. 1. It consisted of an acrylic circular column (1), with an inner diameter of 50 mm and a height of 300 mm. A 0.5 HP centrifugal pump (6) was used for pumping the refractive index-matched liquid. A calming section packed (4) with 34 mm glass beads of 0.3 m height was provided to homogenize the liquid ow before it reached the liquid distributor. The distributor was a perforated plate (3) containing 128 holes of 2 mm diameter on a triangular pitch of 3.1 mm. The circular test section of the uidized bed was encased with a square tank (2) which was lled with a refractive index-matched liquid to allow for Particle Image Velocimetry (PIV) measurements within the bed. 2.2. Solidliquid refractive index matching PIV measurement within opaque uidized beds is only possible at solids concentrations of only a few percent by volume. However, if the refractive index of the particle is matched1 to that of the uid then it is possible to take reliable measurements at solids concentrations greater than 50% by volume. In this study, different mixtures of organic liquid and inorganic salts were screened for matching the refractive index of 1.47 of high precision borosilicate glass beads (dp 3 mm and rp 2230 kg m 3). As indicated in Table 1, from the screening process it was found that Solution 12 (68% turpentine and 32% tetra-hydronaphthalene) provided the
1 An excellent review on the subject of RI matching has been written by Wiederseiner et al. (2011).

best match for the RI of the borosilicate glass beads as well as being compatible with the acrylic column. Parafn oil, with very similar RI to that of the solution and glass beads, was also added in small quantities to change the viscosity of the solution and still maintain the required RI. 2.3. PIV measurements A schematic of the TSI PIV set-up is shown in Fig. 2. The laser source was provided by a pulsed Nd:YAG laser that had a pulse duration of 6 ns and was synchronized with the camera. The time difference between the two laser pulses was optimized based on Nyquist criteria. The optics included a combination of cylindrical and spherical lenses that produced a thin laser sheet of 0.1 mm thickness. Images were captured at a frequency of 2 Hz using a high-resolution 4 M CCD camera (2048 2048 pixels) placed perpendicular to the laser sheet. The refractive index matched liquid was seeded with silver-coated hollow glass particles with a mean diameter equal to 20 mm. The interrogation area was set at 50 100 mm (64 64 pixels, with a 50% overlap) resulting in approximately 1500 vectors per image. Post-processing of the captured raw PIV images was undertaken to determine the velocity vectors. Out-of-plane motion of the seeding particles and strong local velocity gradients caused some spurious velocity vectors. Median ltering, with a threshold value 1.5 times the median of surrounding vectors, was applied to lter the high spurious vectors. A signal-to-noise ratio of 4 was applied to lter the low spurious vectors. Parameters like time

R.K. Reddy et al. / Chemical Engineering Science 92 (2013) 112

Table 1 Screening of refractive index matching liquid for PIV experiments. ID 1 2 3 4 5 6 7 8 9 10 11 12 Liquid 90% W/W glycerin10% W/W water Sodium idiode solution (55% W/W) Parafn oil light Potassium thiocyanate solution (42% W/W) Ammonium thiocyanate solution (45% W/W) P-cyamene Turpentine Benzene or methyl benzoate with turpentine 30% W/W naphthalene70% W/W turpentine Turpentinechloro-naphthalene Turpentinebenzyl alcohol 68% Turpentine32% Tetra hydro naphthalene RI25 1C 1.45 1.475 1.465 1.46 1.47 1.465 1.44 1.47 1.465 1.465 1.47 1.467 Remax 2 210 23 200 200 545 850 900 750 850 850 760 Remarks RI not exactly matched Highly corrosive For low ReN Harmful and corrosive Harmful and corrosive Specialty chemical RI not matched Not compatible with the acrylic column (column was cracked) Corrosive and not stable, turned in to pale yellow Chlorine compounds are not compatible with the acrylic column Not compatible with the acrylic column (column was cracked) RI has been matched and compatible with the acrylic column

Table 2 PIV experimental conditions. PIV exp no. Parafn light oil % W/Wa 85 63 48 34 18 6.5 0.0
a

minimize light contamination from external sources. Additional information2 pertaining to the PIV measurements can be found in Deshpande et al. (2009, 2010) and Sathe et al. (2010).
rL
(kg m 3) RI () VSN (m s 1)

mL
(Pa s)

eL range
()

ReN ()

1 2 3 4 5 6 7

0.01010 0.00420 0.00280 0.00193 0.00165 0.00142 0.00120

855 870 880 890 900 910 915

1.462 1.471 1.468 1.469 1.466 1.470 1.470

0.199 0.266 0.298 0.309 0.317 0.325 0.333

0.520.71 0.460.75 0.560.72 0.530.73 0.510.73 0.460.66 0.490.75

51 164 276 437 525 625 759

3. Direct numerical simulation 3.1. Simulation set-up The physical properties of particles and liquid used to simulate different particle Reynolds numbers within a uidized bed are the same as those used in Reddy et al. (2010b) and are listed in Table 3. Direct numerical modelling was undertaken for 1, 9, 27, 100, 180 and 245 particles contained. For each system, corresponding to a desired particle hold-up, the particles were initially arranged in a regular lattice. At time, t 0, the particles were allowed to settle through a rectangular domain with width and depth of 10 and 20 particle diameters, respectively. A moving reference frame method was employed, resulting in the rectangular domain to move downward with a velocity equal to the average velocity of particle ensemble. At steady state, this situation is comparable to a uidized bed where the heavier particles are suspended by the upward velocity of the liquid. The comparison between the particle and liquid velocity reference conditions is shown in Fig. 3. For the uidized bed (A), the liquid rises upwards with particle settling velocity; whilst for the DNS domain (B) the volume averaged liquid velocity is zero and the particle is moving down with its settling velocity. The domain is sliding downward with the same velocity as the particle. In both cases, the relative velocity of particles with respect to the walls of the domain is zero, and the relative velocity of the liquid with respect to the domain walls is equal to settling velocity of the particles, in the upward direction. The moving reference frame has two advantages for this application. First, very long settling times can be simulated by using a domain with nite dimensions making the simulations computationally affordable; and second, it is a relatively straightforward to establish a uidized bed with a given particle hold-up. The initial conditions for the DNS are summarized in Table 4. The side faces of the box have been treated as wall boundary condition. The top face of the box was treated as a pressure outlet, whilst the bottom face was treated as the velocity inlet with zero velocity magnitude. The initial particle positions are shown in Fig. 4.
2 For a more general discussion on the topic see Tokuhiro et al. (1998), Deen et al. (1999), Lindken and Merzkirch (1999, 2002) and Broder and Sommerfeld (2003).

Added to liquid 12 listed in Table 1.

Synchronizer

Camera
Laser

Solid-liquid fluidized bed

Pump

Refractive index matching liquid

Fig. 2. Schematic diagram of PIV setup.

difference between laser pulses, light sheet thickness and seeding density were optimized so that spurious vectors remained below 2%. The PIV measurements were performed in darkness to

R.K. Reddy et al. / Chemical Engineering Science 92 (2013) 112

Table 3 DNS physical properties. ReN () dP (mm) 1 15 1120 900 0.5700 5 15 1120 900 0.2210 10 15 1120 900 0.1430 30 15 1120 900 0.0680 65 15 1120 900 0.0390 100 15 1120 900 0.0282 200 15 1120 900 0.0166

rp (kg m 3) rL (kg m 3) mL (Pa s)

VL

VL=0

in the last time step. Finally, the solver is parallelized using the PETSc libraries (Jin et al., 2008, 2009) on a shared memory system.

H VS=

Control volume stationary, Liquid moving up with velocity VL

H VS

Liquid stationary, Control volume moving down with velocity VS

4. Results and discussion 4.1. PIV measurements Seven experiments were performed using the refractive index matched solidliquid uidized bed. A summary of the experimental conditions is given in Table 2. Two sample raw images at ReN of 625 and supercial liquid velocities of 0.046 and 0.116 m/s are shown in Fig. 5. The post-processing of instantaneous velocity ow eld at ReN of 625 and supercial liquid velocity of 0.046 m s 1 is shown in Fig. 6. The radial variation of the average axial velocity, vL, at supercial liquid velocities of 0.046 and 0.116 m s 1is shown in Fig. 7. It can be observed that these velocity proles are almost at in the radial direction. It can be observed that at ReN of 625 and at supercial liquid velocity, VL, of 0.047 m s 1 the axial turbulent intensity was on average 69% and the radial turbulent intensity was around 42% (Fig. 8A). An increase in the VL from 0.047 to 0.116 m s 1, results in a decrease in the solid volume fraction of the bed from 0.54 to 0.355. Therefore the turbulent intensity (ratio of root mean square velocity to the average velocity) also decreases to about 47% in the axial direction and around 28% in radial direction (Fig. 8B). It has been observed that at ReN 437 and at VL of 0.047 m s 1 the axial turbulent intensity is 59% and radial turbulent intensity is around 36%. In order to examine the effect of Reynolds number on the turbulent intensity, the Reynolds number has been varied over the range 51759. The variation of average axial and radial turbulent intensities with respect to Reynolds number is shown in Fig. 9. At the lowest Reynolds number of 51 it was observed that the axial turbulent intensity is only 7% whilst the radial turbulent intensity is around 2%. When the Reynolds number is increased from 51 to 164 it was observed that the axial and radial turbulent intensities increase to 19% and 8%, respectively. Further increase in Reynolds number from 164 to 759 led to an axial turbulent intensity of 75% and a corresponding radial turbulent intensity of 47%. In the turbulent regime, at ReN 4500 (Clift et al., 1978; Joshi, 1983), it was found that the axial turbulent intensity is 1.65 times that of the radial turbulent intensity. This value is lower than the value of 2.5 reported by Handely et al. (1966). The image capture and processing technique employed here, utilizing a laser light sheet to clearly illuminate the particle boundary, is arguably more precise than the photographic measurement of Handely et al. (1966). Based on the limited number of experimental measurements, the following simple relationships are proposed for the relationships between the turbulence intensity in the axial, v0 , and radial, u0 , directions and the solid phase holdup, AS, and slip velocity, VS: v0 1:29 A S VS u0 0:78A S VS 7

VL
Fig. 3. Schematic of uidized bed and moving reference frame DNS simulation framework.

3.2. Fictitious domain formulation The DNS simulations were carried out using the ctitious domain method described previously (see Diaz-Goano et al., 2003; Veeramani et al., 2007). Briey, the uid and particles occupy domains O1 and O2, respectively, and the relevant equations of continuity and motion are applied to each domain as appropriate. In the ctitious domain approach the equations for the uid are extended into the particle domain, such that O (O1UO2), resulting in savings in computational time as the liquid domain no longer needs to be re-meshed each time there is particle motion. The relevant equations3 are
r2,i r1 Du1 2 1 rp1 Re r u1 r1 GF , Dt

rUu1 0 inO

and DU 1 1 Vi Dt where ( G and ( F and ( ^ F


1 Du1 Re r2 u1 rp1 Dt

Z
O2,i

FdO

1 Fr eg

inO2,i , inO1

i 1,:::,n

r1 1 Fr eg r2,i r1

^ F

inO2,i , inO1

i 1,:::,n

inO2,i , inO1

i 1,:::,n

where F is the interaction force. For collision of a particle with the plane wall, the interaction force can be replaced by the lubrication force given by tenCate et al. (2002):   ( ^ 6pr i U ? m1 r^i ri if h o h h ^W h Fi 6 0 otherwise An operator splitting procedure is applied to the discretization process, with a rigid body constraint being applied to the particles
3

See Veeramani et al. (2007) for denition of terms.

R.K. Reddy et al. / Chemical Engineering Science 92 (2013) 112

Table 4 DNS initial conditions. Number of particles dP (mm) rp (kg m 3) Ln () Comp. domain dimension (mm) (mm) (mm) Number of nodes () 1 15 1120 8 16 8 1.2 106 9 15 1120 1.2 8 16 8 1.9 106 27 15 1120 1.2 8 16 8 2.6 106 100 15 1120 1.2 8 16 8 4.1 106 180 15 1120 1.2 8 16 8 5.2 106 245 15 1120 1.2 8 16 8 6.5 106

Fig. 4. Initial particle positions for (A) 1, (B) 9, (C) 27, (D) 100, (E) 245 particles [see Fig. 14A for orientation for 180 particles].

Fig. 5. Raw images at ReN 625 for VL (A) 0.046 and (B) 0.116 m s 1.

The radial variation of turbulent kinetic energy at different Reynolds numbers is shown in Fig. 10. It can be seen that as Reynolds number increases from 51 to 759, the turbulent kinetic energy also increases from 2 10 5 to 4 10 3 m2 s 2. Moreover, for each Reynolds number there is practically no

variation in the turbulent kinetic energy prole along the radial direction, which is similar to that for the turbulent intensity proles. Both of these observations are further validations of the homogenous nature of turbulence in the solidliquid uidized bed.

R.K. Reddy et al. / Chemical Engineering Science 92 (2013) 112

Axial velocity (m s 1)

80 70 60 50 40 30 20 10

Fig. 6. Instantaneous axial velocity ow eld at ReN 625 and VL 0.047 m s 1.

0 0 0.2 0.4 0.6 0.8 1

0.14

60

0.12

50
0.1

40
0.08

2
30

0.06

0.04

20

0.02

1
10

0 0 0.2 0.4 0.6 0.8 1

0 0
Fig. 7. Average axial velocity vs radial position at ReN 625 and VL (1) 0.047 and (2) 0.116 m s 1. Fig. 8. Turbulent intensities vs axial and radial position at ReN 625 and VL (1) 0.047 and (2) 0.116 m s 1.

0.2

0.4

0.6

0.8

4.2. Direct numerical simulations 4.2.1. Wake structure Fig. 11 shows the wake structure for a particle both in isolation and with other eight particles present. It can be observed that at the wake structure of a single particle at ReN 200 in (A) an innite medium; and (B) surrounded by eight other particles at centre-to-centre distance of 1.2 dp. It can be seen that whilst the separation angle was similar for both particles the wake length was reduced by almost 33% due to the presence of other particles. For a single particle in an innite medium the axisymmetric toroidal vortex of the wake has been observed up to ReN 210. This same behaviour is exhibited for the isolated particle shown in Fig. 11A, which is not surprising given that ReN 200. For the case of the nine particles this axisymmetry in the wake is broken as shown in Fig. 11B. The reason for this early instability in the wake is attributed to the asymmetric velocity gradients around the particle, resulting in centrifugal acceleration in the core of the toroidal vortex and an increase in the azimuthal velocity. The axisymmetric toroidal vortex eventually breaks into two counter rotating vortices. Additional DNS simulations were undertaken to investigate the inuence of particle spacing on the length of the wake, and particularly to determine the minimum spacing required so that the wake is no longer inuenced by the surrounding particles. The results (for ReN 200) are shown in Fig. 12 for dimensionless wake length, W*, as a function of dimensionless particle spacing between particles, L*. It can be seen for very close particle spacing then W* o1; and it steadily increases to unity at L* equal to 6. Therefore, it can be concluded that the effect of the surrounding particles can be neglected when the centre-to-centre particle spacing is more than 6dp.

R.K. Reddy et al. / Chemical Engineering Science 92 (2013) 112

80
TURBULENT INTENSITY, uiRMS / u2 x 100, (-)

70 60 50 40 30 20 10 0 0 200 400 600 800


Fig. 11. Wake of the sphere at ReN 200 [(A) settling in the innite medium; (B) Settling in the presence of eight other surrounding particles].

REYNOLDS NUMBER, Re (-)


Fig. 9. Average () axial and radial (m) turbulent intensities vs ReN at VL 0.047 m s 1.

0.01

DIMENSIONLESS WAKE LENGTH (-)

0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2

TURBULENT KINETIC ENERGY, k, (m2/s2)

0.001

0.0001

0.00001 0 0.2 0.4 0.6 0.8 1

NORMALIZED RADIAL DISTANCE, r/R, (-)


Fig. 10. Turbulent kinetic energy radial prole vs ReN at VL 0.047 m s 1 [ReN 51(D), 164(J), 276(m), 437(), 795()].

Fig. 12. Wake length vs particle spacing for ReN 200.

4.2.2. Normalized settling velocity The simulated normalized settling velocities (shown as solid lines) for 1, 2, 9, 27, 100, 180 and 245 particles are plotted as a function particle Reynolds number in Fig. 13. It can be seen that the time averaged settling velocity of the particle in the presence of other particles decreases with an increase in the number of particles surrounding it. This is in line with the hindrance effect observed in particles settling in swarm. In the presence of eight other particles, at ReN 1, the normalized average settling velocity, VS, is 5.5% less than that for a single particle, VSN, settling in an innite liquid. For a system of 27 particles the average settling velocity was found to be 0.606, which is 39.4% lower than VSN; whilst for 100 particles VS is only 27% of VSN. The DNS results show clearly that the particle settling velocity decreases with an increasing number of particles, or solids volume fraction, AS, within the computational domain. Given that a decrease in

particle velocity is a consequence of an increase in the drag coefcient, the simulation result is consistent with the ndings of Joshi (1983) and Pandit and Joshi (1998) who used an energy balance approach to derive the following relationship for the particle drag coefcient, CD, as a function of AL: CD A 4:8 L C D1 9

In creeping ow the drag coefcient in an assemblage of particle is increased due to (i) increased true uid velocity within the interstices between the particles, (ii) increased velocity gradients resulting from more zero slip boundary condition surfaces; and (iii) increased length of the uid ow through the assemblage of particles. In turbulent ow the same conditions prevail, with enhanced momentum transfer, and hence higher drag as the solids concentration is increased.

10

R.K. Reddy et al. / Chemical Engineering Science 92 (2013) 112

incorporation of a collision model that captures all of the dynamics is an area of on-going DNS research. 4.2.4. Energy dissipation rate The computed energy dissipation rate, e, was calculated from the three components of instantaneous velocity and nine velocity gradients at 12 points in the computational domain using the expression:       @u1 2 @u2 2 @u3 2 e2 2 2 @x1 @x2 @x3  2  2         @u1 @u1 @u2 2 @u2 2 @u3 2 @u3 2 @x2 @x3 @x1 @x3 @x1 @x2       @u1 @u2 @u2 @u3 @u3 @u01 2 2 2 10 @x2 @x1 @x3 @x2 @x1 @x3 The computed averaged energy dissipation rate for 245 particles at ReN 51, with up to 3500 time steps was found to be 0.303 m2 s 3. This value compares favourably with the experimental value of 0.36 m2 s 3 obtained by a volume-averaged energy balance over the refractive index matched uidized bed used in this study4 at the same ReN 51. The local energy dissipation rate computed from the PIV measurements at the point equivalent of 0.6, 8.7, 0.6 in the computational domain is shown in Fig. 15 as a function of normalized time, t*. Two peeks can be clearly seen at t*34.8 and 43.5 during the measurement time. The rst peak of approximately 4.0 m2 s 3 corresponded to the approach of a particle to the measurement position, whilst the second peak of 5.5 m2 s 3 corresponded to the departure of the particle. The period between these two peaks corresponded to the time when the particle occupied the sampling volume and hence a no liquid velocity vectors were measured and a zero local energy dissipation rate was recorded. The peaks in the local energy dissipation rate at the surface of the particle represent an increase of between 10 and 18 times than that of the average energy dissipation rate of 0.36 m2 s-3. The ow patterns (including the energy dissipation rate) in the vicinity of the particle interface and in the bulk have different roles to play in terms of industrial operations. The ow pattern near the particle surface governs the heat and mass transfer characteristics at the particleuid interface, whereas the ow pattern in the bulk governs the solid and liquid phase dispersions. Typically, high values of particleuid heat and mass transfer coefcients are desired along with low levels of particle and uid phase dispersion so that plug ow can be achieved. These two characteristics are achieved by high and low levels of energy dissipation rate, respectively. This behaviour is reected in Fig. 15 as a particle transits through the sampling location, and for this reason it is perhaps not surprising that uidized beds are so widely used for heat and mass transfer applications. However, there are some limitations. First, the low dissipation rate in the bulk results in low value of mass and heat transfer coefcients at the container wall. Second, there may be some practical cases where complete bulk mixing is desired in the bulk. For achieving these process objectives an optimum selection of design (including column diameter and height) and the operating parameters (such as particle size and density, supercial liquid velocity, liquid viscosity and density) is required.

Fig. 13. Average velocity of the particles: 1, single particle; 2, nine particles; 3, 27 particles; 4, 100 particles; 5, 180 particles; 6, 245 particles, DNS (solid line), , Richardson and Zaki (1954) For 1 o Re1 o 200 n 4:45 18 dp =D Re0:1 1 For 200 o Re1 o 500 n 4:45Re0:1 . 1

Also shown in Fig. 13, is the corresponding normalized settling velocity predictions using the correlation of RichardsonZaki (1954) where the liquid void fraction, AL, used for the RZ calculation was that obtained from the DNS for the different number of particles. It can be seen that the DNS simulations are in good agreement (maximum deviation of 24%) with the RZ correlation. Interestingly, the RZ correlation is based on a very large number of experimental studies, typically for systems with a homogeneous dispersion of a very large number of particles (at least 2000). Whilst the current DNS simulations are for a relatively small number particles. It can be seen from Fig. 13 that the agreement is increasing with increasing number of particles where the inuence of the wall on the wake of an individual particle in the bulk assemblage is less. Ideally, more particles should be included in the simulations. For this study, however, a 64 IBM Power4 processors having 256 GB RAM was required to satisfactorily resolve the computational domain of 6.5 million nodes for the 245 particles at ReN 200. Further increase in either the numbers of particles or Reynolds numbers will require much bigger domains. To do this, the code will need to be fully parallelized using domain decomposition technique in order to perform simulations within a reasonable time frame.

4.2.3. Particleparticle and particlewall collisions The time sequence at 0, 3, 6, 9 and 12 s for the sedimentation of 180 particles at ReN 200 is shown in Fig. 14. The images show that at different time intervals different particles come in contact with each other, which is qualitatively in agreement with experimental observations. For the simulations in this study a simple lubrication model was used to account for both the particle particle and particlewall collisions. The lubrication model is really only valid for low Reynolds numbers, and at best an approximation at higher Reynolds numbers. Moreover, the model does not account for the experimentally observed clusters, especially at higher solids concentrations, where two or more particles remain in contact with each other for some time. The

5. Conclusions Flow visualization experiments were performed using particle image velocimetry and refractive index matching of the solid and
4

PIV experiment 1 as listed in Table 2.

R.K. Reddy et al. / Chemical Engineering Science 92 (2013) 112

11

Fig. 14. Sedimentation of 180 particles at different time intervals for ReN 200 [(A) t 0; (B) t 3; (C) t 6; (D) t 9; (E) t 12].

particle surface whilst mixing of the liquid and solid phases will be controlled by the dissipation rate in the bulk.

Nomenclature uctuating concentration (kmol m 3) drag coefcient of a single particle in innite medium () CD drag coefcient of a single particle in the presence of neighbouring particles () dP diameter of the particle (m) eg unit vector in the direction of gravity Fr Froude number (V2 =gdp ) () S1 F dimensionless redenition of interaction force in the most convenient form () G representation of gravitational force in Eq. (6) ^ h distance between particle and wall or particle and particle (m) h grid size (m) L* dimensionless particle spacing between particles () ^ pL extension of the pL to the entire domain O () ReN Reynolds number based on the particle (dp VsNrL/mL) () ri dimensionless radius of ith particle () time (s) t t* normalized time () U? velocity component perpendicular to the wall or particle (m s 1) ui Velocity of uid in ith direction (m s 1) ^L u dimensionless uid velocity () ^ uL extension of the uL to the entire domain O u0 radial uctuation velocity (m s 1) v0 axial uctuation velocity (m s 1) Vi volume of the ith particle VL supercial liquid velocity for uidization (or) hindered settling velocity of the particle in sedimentation (m s 1) VS1 terminal settling velocity (m s 1) Vs interstitial velocity (m s 1) W wake length (m) (Wn W/dp) before the dimension () c0 CDN

Fig. 15. Local energy dissipation rate vs time for ReN 51 [measurement taken at the point (0.6, 8.7, 0.6)].

liquid phases to understand the characteristics of turbulence in a uidized bed. For the Reynolds number in the range of 51759 the experimental measurements revealed that the turbulence intensity was constant in both the radial and axial directions, thus establishing the homogenous nature system. Complementary DNS simulations provided increased spatial resolution of the velocity eld that could be obtained by the PIV measurements. The computational domain, moving in the downward direction with velocity equal to the averaged velocity of particles in the swarm, comprised 1, 9, 27, 100, 180 and 245 particles. For each case, the wake of individual particle was observed to attenuate with increase in the volume fraction of particles. The averaged particle slip velocity decreased with increase in the number of surrounding particles, and compared reasonably with the Richardson and Zaki (1954) correlation. Finally, the local energy dissipation rate was computed from the DNS simulation, and for ReN 51, the energy dissipation rate near the surface of the particle was found to be approximately 18 times the volume averaged energy dissipation in the uidized bed. Such a variation in energy dissipation rate distribution would need to be taken into consideration when designing uidized beds given that heat and mass transfer will be controlled by what is occurring at the

Greek symbols AL AS volume fraction of uid volume fraction of solid

12

R.K. Reddy et al. / Chemical Engineering Science 92 (2013) 112

rp rS,i rL
OL O OS

mL e

density of solid particle (kg m 3) density of ith solid particle (kg m 3) density of uid (kg m 3) uid domain entire computational domain solid domain molecular viscosity of uid (kg m 1 s 1) energy dissipation rate (m2/s 3)

Subscripts i L S particle species liquid phase solid phase

References
Broder, D., Sommerfeld, M., 2003. Combined PIV/PTV-measurements for the analysis of bubble interactions and coalescence in a turbulent ow. Can. J. Chem. Eng. 81, 756763. Chen, R.C., Kadambi, J.R., 1990. LDV measurements of solidliquid ows. Particulate Sci. Technol. 8, 97109. Chen, R.C., Fan, L.S., 1992. Particle image velocimetry for characterizing the ow structure in three-dimensional gasliquidsolid uidized beds. Chem. Eng. Sci. 47, 36153622. Clift, R., Grace, J.R., Weber, M.E., 1978. Bubbles, Drops and Particles. Academic Press, New York. Deen, N.G., Annaland, M.v.S., Kuipers, J.A.M., 2009. Direct numerical simulation of complex multi-uid ows using a combined front tracking and immersed boundary method. Chem. Eng. Sci. 64, 21862201. Deshpande, S.S., Sathe, M.J., Joshi, J.B., 2009. Evaluation of local turbulent energy dissipation rate using PIV in jet loop reactor. Ind. Eng. Chem. Res. 48, 50465057. Deshpande, S.S., Tabib, M.V., Joshi, J.B., Ravi Kumar, V., Kulkarni, B.D., 2010. Analysis of ow structures and energy spectra in chemical process equipment. J. Turbulence 11, 139. Diaz-Goano, C., Miniv, P.D., Nandakumar, K., 2003. A ctitious domain/nite element method for particular ows. J. Comput. Phys. 192, 105123. Dijkhuizen, W., Bokkers, G.A., Deen, N.G., Van Sint Annaland, M., Kuipers, J.A.M., 2007. Extension of PIV for measuring granular temperature eld in dense uidized beds. AIChE J. 53, 108118. Garside, J., A1-Dibouni, M.R., 1977. Velocity-voidage relationships for uidization and sedimentation in solidliquid systems. Ind. Eng. Chem. Process Des. Dev. 16, 206214. Gevrin, F., Masbernat, O., Simonin, O., 2008. Granular pressure and particle velocity uctuations prediction in liquid uidized beds. Chem. Eng. Sci. 63, 24502464. Gidaspow, D., 1994. Multiphase Flow and Fluidization: Continuum and Kinetic Theory Descriptions. Academic Press, San Diego. Haam, S.J., Brodkey, R.S., Fort, I., Klaboch, L., Placnik, M., Vanecek, V., 2000. Laser doppler anemometry measurements in an index of refraction matched column in the presence of dispersed beads. Int. J. Multiphase Flow 26, 14011418. Handely, D., Doraisamy, A., Butcher, K.L., Franklin, N.L., 1966. A study of the liquid and particle mechanics in liquid uidized beds. Trans. Inst. Chem. Eng. 44, T260T273. Hanratty, T.J., Bandukwala, A., 1957. Fluidization and sedimentation of spherical particles. AIChE J. 3, 293296. Hernandez-Jimenez, F., Sanchez-Delgado, S., Gomez-Garca, A., Acosta-Iborra, A., 2011. Comparison between two-uid model simulations and particle image analysis & velocimetry (PIV) results for a 2D gassolid uidized bed. Chem. Eng. Sci. 66, 37533772. Jin, S., Veeramani, C., Minev, P. and Nandakumar, K., 2008. A Parallel Algorithm for the Direct Numerical Simulation of 3D Inertial Particle Sedimentation. In: 16th

Annual Conference of the CFD Society of Canada. Saskatoon, Canada, 810 June 2008. Jin, S., Minev, P., Nandakumar, K., 2009. A scalable parallel algorithm for the direct numerical simulation of 3D incompressible particulate ow. Int. J. Comput. Fluid Dyn. 23, 427437. Joshi, J.B., 1983. Solidliquid uidized beds: some design aspects. Chem. Eng. Res. Des. 61, 143161. Joshi, J.B., Sharma, M.M., Shah, Y.T., Singh, C.P.P., Ally, M., Klinzing, G.E., 1980. Heat transfer in multiphase contactors. Chem. Eng. Commun. 6, 257271. Joshi, J.B., Shah., Y.T., 1981. Hydrodynamics and mixing models for bubble column reactors. Chem. Eng. Commun. 11, 165199. Joshi, J.B., Ranade, V.V., 2003. Computational uid dynamics for designing process equipment: Expectations, current status and path forward. Ind. Eng. Chem. Res. 42, 11151128. Kashyap, M., Chalermsinsuwan, B., Gidaspow, D., 2011. Measuring turbulence in a circulating uidized bed using PIV techniques. Particuology 9, 572588. Lindken, R., Merzkirch, W., 1999. Velocity measurement for liquid and gaseous phase for a system of bubbles rising in water. Exp. Fluids, Supplement: s194s201. Lindken, R., Merzkirch, W., 2002. A novel PIV technique for measurements in multiphase ows and its application to two-phase bubbly ows. Exp. Fluids 33, 814825. Muller, C.R., Hartung, G., Hult, J., Dennis, J.S., Kaminski, C.F., 2009. Laser diagnostic investigation of the bubble eruption patterns in the freeboard of uidized beds: Simultaneous acetone PLIF and stereoscopic PIV measurements. AIChE J. 55, 13691382. Murthy, B.N., Ghadge, R.S., Joshi, J.B., 2007. CFD simulations of gasliquidsolid stirred reactor: Prediction of critical impeller speed for solid suspension. Chem. Eng. Sci. 62, 71847195. Pandit, A.B., Joshi, J.B., 1998. Pressure drop in packed, expanded and uidized beds, packed columns and static mixers a unied approach. Rev. Chem. Eng. 14, 321371. Reddy, R.K., Joshi, J.B., 2009. CFD modelling of solidliquid uidized beds of mono and binary particle mixtures. Chem. Eng. Sci. 64, 36413658. Reddy, R.K., Jin, S., Joshi, J.B., Nandakumar, K., Minev, P.D., 2010a. Direct numerical simulation of free falling sphere in creeping ow. Int. J. Comput. Fluid Dyn. 24, 109120. Reddy, R.K., Joshi, J.B., Nandakumar, K., Minev, P.D., 2010b. Direct numerical simulations of wake generated by a freely falling sphere. Chem. Eng. Sci. 65, 21592171. Richardson, J.F., Zaki, W.N., 1954. Sedimentation and uidization: part I. Trans. Inst. Chem. Eng. 32, 3553. Sathe, M.J., Thaker, I.H., Strand, T.E., Joshi, J.B., 2010. Advanced PIV/LIF and shadowgraphy system to visualize ow structure in two-phase bubbly ows. Chem. Eng. Sci. 65, 24312442. Shi, H., 2007. Experimental research of ow structure in a gassolid circulating uidized bed riser by PIV. J. Hydrodyn. 19, 712719. Syamlal, M., OBrien, T.J., 1989. Computer simulation of bubbles in a uidized bed. AIChE Symp. Ser. 85, 2231. tenCate, A., Nieuwstad, C.H., Derksen, J.J., van den Akker, H.E.A., 2002. Particle imaging velocimetry experiments and lattice-Boltzmann simulations on a single sphere settling under gravity. Phys. Fluids 1411, 40124025. Thakre, S.S., Joshi, J.B., 1999. CFD simulation of ow in bubble column reactors importance of drag force formulation. Chem. Eng. Sci. 54, 50555060. Tenneti, S., Garg, R., Subramaniam, S., 2011. Drag law for monodisperse gassolid systems using particle-resolved direct numerical simulation of ow past xed assemblies of spheres. Int. J. Multiphase Flow 37, 10721092. Uhlmann, M., Pinelli, A., 2009, Direct numerical simulation of vertical particulate channel ow in the turbulent regime. In: Proceedings of the 20th International Conference on Fluidized Bed Combustion, pp. 8396. Van Buijtenen, M.S., Buist, K., Deen, N.G., Kuipers, J.A.M., Leadbeater, T., Parker, D.J., 2011. Numerical and experimental study on spout elevation in spout-uidized beds. AIChE J. 58, 25242535. Veeramani, C., Minev, P.D., Nandakumar, K., 2007. A ctitious domain formulation for ows with rigid particles: a non-Lagrange multiplier version. J. Comput. Phys. 224, 867879. Wiederseiner, S., Andreini, N., Epely-Chauvin, G., Ancey, C., 2011. Refractive-index and density matching in concentrated particle suspensions: a review. Exp. Fluids 50, 11831206. Xu, Y., Subramaniam, S., 2010. Effect of particle clusters on carrier ow turbulence: a direct numerical simulation study. Flow, Turbulence Combust. 85, 735761. Zivkovic, V., Biggs, M.J., Glass, D., Pagliai, P., Buts, A., 2009. Granular temperature in a liquid uidized bed as revealed by diffusing wave spectroscopy. Chem. Eng. Sci. 64, 11021110.

You might also like