Experimental Thermal and Fluid Science 28 (2004) 407–421 www.elsevier.com/locate/etfs

CFD simulation of sparger design and height to diameter ratio on gas hold-up profiles in bubble column reactors M.T. Dhotre, K. Ekambara, J.B. Joshi

*

Institute of Chemical Technology, University of Mumbai, Matunga, Mumbai 400 019, India Received 10 August 2002; accepted 15 June 2003

Abstract In the present work, a two-dimensional CFD model has been developed for the prediction of flow pattern in bubble column reactors. The model has been validated using available experimental data and extended to simulate the effect of the sparger design and height to diameter ratio on radial gas hold-up profiles. The predictions were compared with experimental measurements for a 0.385 m i.d. bubble column. The complete energy balance was established in all the cases. The simulations were carried out for three different gas–liquid systems (air–water, air–aqueous solution of butanol and air–aqueous solution of carboxyl methyl cellulose). A comparison has been presented between the predicted and the experimental data over a wide range of superficial gas velocity and for three gas–liquid systems. In all these cases, the CFD model has been found to predict the variation of hold-up profiles with respect to the column height and the sparger design.  2003 Elsevier Inc. All rights reserved. Keywords: Computational fluid dynamics; Bubble column; Flow patterns; Sparger design; Height to diameter ratio; Gas–liquid dispersions

1. Introduction Bubble columns are widely used in industry because of their simple construction and operation. Important applications include oxidation, hydrogenation, halogenation, hydrohalogenation, ammonolysis, hydroformylation, Fischer–Tropsch reaction, ozonolysis, carbonylation, carboxylation, alkylation, fermentation, waste water treatment, hydrometallurgical operations, steel ladle stirring, column flotation, etc. The fractional gas hold-up (G ) is an important parameter in the design and scale-up of bubble column reactors. It has several direct and indirect influences on the column performance. The direct and obvious effect is on the column volume. This is because the fraction of the volume occupied by the gas and the respective phase volume becomes important depending upon the phase in which the rate controlling step takes place. The indirect influences are far reaching. The spatial variation of G , gives *

Corresponding author. Tel.: +91-22-2414-5616; fax: +91-22-24145614. E-mail address: [email protected] (J.B. Joshi). 0894-1777/$ - see front matter  2003 Elsevier Inc. All rights reserved. doi:10.1016/j.expthermflusci.2003.06.001

rise to pressure variation and eventually results in intense liquid phase motion. These secondary motions govern the rate of mixing, heat transfer and mass transfer. In bubble columns, the gas phase exists as a dispersed bubble phase in a continuous liquid phase. The gas phase moves in one of the two characteristic regimes. The two regimes are: homogeneous and heterogeneous. The homogeneous regime occurs at relatively low superficial gas velocities (less than about 50–80 mm/s). This regime is characterized by almost uniformly sized bubbles and the concentration of bubbles is also uniform, particularly in the transverse direction. The heterogeneous regime occurs at relatively high superficial gas velocities. This regime is characterized by the presence of a radial hold-up profile as against a flat profile in the homogeneous regime. In the heterogeneous regime, the role of sparger design diminishes depending upon the column height. In fact, the total column height can be divided into two regions: the sparger region and the bulk region. The size of the bubbles formed at the sparger (primary bubble size dBP ) depends upon the sparger design, the local energy dissipation rate and the surface active contaminants.

408

M.T. Dhotre et al. / Experimental Thermal and Fluid Science 28 (2004) 407–421

Nomenclature Air–alcohol air–aqueous solution of butanol Air–CMC air–aqueous solution of carboxyl methyl cellulose CB interface energy transfer factor CD drag force coefficient CL lift force coefficient CV virtual mass force coefficient C0 , C1 drift flux constants Ce1 model parameter in turbulent dissipation energy equation (¼ 1.44) Ce2 model parameter in turbulent dissipation energy equation (¼ 1.92) D diameter of the column (m) E rate of energy released by all the bubble in the column (W) ES fraction of dissipation energy (W) FDR frictional force in the radial direction per unit volume of dispersion (N/m3 ) FDZ frictional force in the axial direction per unit volume of dispersion (N/m3 ) FL lift force per unit volume of dispersion (N/ m3 ) FVZ axial virtual mass force per unit volume of dispersion (N/m3 ) FVR radial virtual mass force per unit volume of dispersion (N/m3 ) g gravitational (m/s2 )   constant v 2 ou 2  ov  ovL 2 G ¼ lt;L 2 or þ rL þ ozL þ ozL ou 2  þ orL H HD k P PB r R u

height of the bubble column (m) height of the gas dispersion (m) turbulent kinetic energy per unit mass (m2 /s2 ) pressure (N/m2 ) interphase transfer of energy term. CB ½FDr VSr þ FDz VS  radial distance (m) column radius (m) instantaneous axial velocity (m/s)

In the sparger region, the bubble size changes with respect to height depending upon the coalescence nature of the liquid phase, the extent of turbulence and the bulk motion. At the end of the sparger region, the bubble attains an equilibrium size (called secondary bubble size, dBS ). The equilibrium is governed by the breaking forces due to bulk motion (turbulent and viscous stresses) and the retaining force due to surface tension. The height of the sparger region depends upon the difference between dBP and dBS , the coalescing nature of the liquid phase and the liquid circulation in the heterogeneous regime. The relative proportion of the sparger region in the total

uG uL u v vG vL v VC VG VL VSr VS z

axial component of gas velocity (m/s) axial component of liquid velocity (m/s) axial time-averaged velocity (m/s) instantaneous radial velocity (m/s) radial component of gas velocity (m/s) radial component of liquid velocity (m/s) radial time-averaged velocity (m/s) circulation velocity (m/s) superficial gas velocity (m/s) superficial liquid velocity (m/s) radial slip velocity between gas and liquid (m/s) axial slip velocity between gas and liquid (m/s) axial distance along the column (m)

Greek symbols CK lK þ lt;K =r/ a, b proportionality constants in Eq. (10)  fractional phase hold-up G fractional gas hold-up L fractional liquid hold-up L time-averaged fractional liquid hold-up e turbulent energy dissipation rate per unit mass (m2 /s3 ) lK molecular viscosity of phase K (Pa s) lt;K turbulent viscosity of phase K (Pa s) m molecular kinematic viscosity of liquid (m2 /s) mt turbulent kinematic viscosity of liquid (m2 /s) q density (kg/m3 ) qG density of the gas (kg/m3 ) qL density of the liquid (kg/m3 ) r/ turbulent Prandtl number for momentum transfer rf turbulent Prandtl number for bubble motion Subscripts G gas phase K phase, K ¼ G: gas phase, K ¼ L: liquid phase L liquid phase

column height decides the effect of HD =D ratio on G . If the sparger region is small, the effect of HD =D ratio on G is small and vice-versa. The fractional gas hold-up in bubble columns has been extensively investigated during the last five decades and more than 200 papers are available in the published literature. As regards to the effect of the HD =D ratio on G , the following observations can be noted from the published literature. (i) In the case of air–water systems, for multipoint spargers (having do < 3 mm), the fractional hold-up is

M.T. Dhotre et al. / Experimental Thermal and Fluid Science 28 (2004) 407–421

(ii)

(iii)

(iv) (v)

(vi)

(vii)

maximum at very low HD =D (say 0.259) and decreases by 15–20% as HD =D increases upto a value of 5. A further increase in the HD =D ratio results into marginal decrease in average fractional holdup. When the hole diameter is larger than 3 mm, the HD =D ratio has nominal influence on average fractional hold-up [1]. For single point spargers, as HD =D increases from 1 to 5, the average hold-up increases by about 50– 100%. Again, a further increase of HD =D upto 7 increases average hold-up marginally and thereafter it remains practically constant [1]. In the case of air–electrolyte systems, the value of average gas hold-up increases with an increase in electrolyte concentration. However, the effect of concentration levels off at a critical concentration. A further increase in the electrolyte concentration does not have any influence on average hold-up [2]. For multipoint as well as single point spargers, the variation of average hold-up with respect to HD =D is slower (10–12% and 10–15%, respectively.) as compared to the variation in the air– water system (20–22% and 50–100%, respectively). The effect of HD =D levels off at a ratio of 5 for air– water system whereas in electrolyte solutions, the ratio extends up to 8 [3]. Similar to electrolytes, aqueous solutions of aliphatic alcohols show similar behavior [4]. In the case of air–carboxyl methyl cellulose (CMC) systems, for multipoint as well as single point spargers, the variation of average hold-up with respect to HD =D is faster as compared to the variation in an air–water system. This behavior is attributed to the coalescing nature of these dispersions [3]. For the single point sparger, the profiles are found to be steep near the sparger and they become flatter with an increase in distance from the sparger. For multipoint spargers, the profiles are relatively flat at the bottom and they become steeper with an increase in the height of dispersion [3]. With an increase in height of dispersion, the profiles undergo a shape change of flatter to steeper and reverse change occurs for single point spargers. However, this type of variation occurs up to HD =D ratio of 5 for air–water and HD =D ratio of 3 for air–CMC systems. Beyond this, the shapes are practically the same irrespective of sparger design for air–water and air–CMC systems, respectively. With a further increase in HD =D, the shape changes were found to be nominal. For air–alcohol, the profile remains flat even after HD =D ¼ 5 without undergoing shape change. In case of air–alcohol system, sparger influence remains up to a much higher HD =D ratio, as compared with air–water and air–CMC systems [3].

409

(viii) In the majority of papers where the sparger design and HD =D ratio were found to be important, the major impact was due to transition from homogeneous to heterogeneous regime and vice-versa. This was principally because of the selection of small diameter columns (<150 mm) and the lower range of superficial gas velocity (100 mm/s). From the foregoing discussion, it is clear that the hold-up profile (and hence DG ) depends upon VG , D, HD , sparger design and the nature of the gas–liquid system. Therefore, the aim of the present modeling work is to understand the influence of the sparger configuration on the fractional gas hold-up. To achieve this purpose, a computational model has been developed, building upon the on-going developments which have already been achieved in the computational fluid dynamics modelling of two phase systems [5]. An attempt has been made to incorporate two different spargers and study the effects of superficial gas velocity, sparger design and height of dispersion on the hold-up profile. 2. Mathematical model For the case of bubbly flow, the equations of continuity and motion for a two-dimensional cylindrical coordinate system can be represented in the following generalized form: 1 o o ðrqvUÞK þ ðquUÞK r or oz     1 o oU o oU rC C ¼ þ þ SU;K ð1Þ r or or K oz oz K where U is the transport variable (for instance, U is one for equation of continuity or it is a velocity component for the respective equation of motion, and U ¼ k and e for the conservation equation for the turbulent kinetic energy and turbulent energy dissipation rate respectively). K denotes the phase [K ¼ G or L] and SU is the source term for the respective dependent variable. The values of U and SU for different transport variables have been given in Table 1. From Table 1, it can be seen that most of the terms are derived from gas and liquid velocities. In addition, in two-phase flows, momentum and energy transfer occurs across the interface. Therefore, the drag force (FDZ and FDR ), virtual mass force (Fvz and Fvr ), and lift force (FLr ) appear in the axial and radial components of the momentum balance (Table 1). Furthermore, the interphase transfer of energy (PB ) appears in the equations for turbulent kinetic energy and turbulent energy dissipation rates. 2.1. Boundary and initial conditions (i) along the axis: axisymmetry in uG , uL , G , vG , vL , k and e;

410

Table 1 Two-dimensional k–e model for bubble column: governing equations The governing equations written in a general form: Conservation of

U

Mass

1

1

1–1

    lt;K oK 1o o lt;K oK r þ r or oz rf oz rf or

Axial velocity momentum

u

1.0

1–1

K

Radial velocity momentum

v

1.0

1–1

K

Turbulent kinetic energy

k

1.0



L ðG þ PB  qL eÞ

Turbulent dissipation energy

e

1.3



e L ðCe1 ðG þ PB Þ  Ce2 qL eÞ k

lt;K ¼ 0:09qk ðk 2 =eÞ;

FVZ ¼ CV G qL

FVR ¼ CV G qL G ¼ lt;L 2

PB ¼ CB ðFDr VSr þ FDZ VS Þ;

FLr ¼ CL G qL ðuG  uL Þ

     oP 1o ov o ou   þ K qK g  FDZ rlt þ lt L  FVZ L þ oz r or oz oz oz K          1o lt o o lt o lt o 1o ou r þ ðrvÞ þ þuK þ r or rf or oz rf oz K oz K rf oz K r or      oP 1o ov o ou     FDR rlt þ lt L  FVR L  FLr L þ or r or oz oz or K            1o lt o o lt o lt o 1o ou v r þ ðrvÞ þ  2lt r2 K þ vK þ r or rf or oz rf oz K oz K rf or K r or

ouL or

D ðuG  uL Þ Dt

FDZ ¼

D ðvG  vL Þ Dt

FDR ¼

ovL or

2 þ

n v o2 L

r

þ

ouL oz

2 !

 þ

ovL oz



þ

ouL or

2 !

G ðqG  qL ÞgðuG  uL ÞjuG  uL j ðuG  uL Þ2 G ðqG  qL ÞgðvG  vL ÞjvG þ vL j ðvG  vL Þ2

M.T. Dhotre et al. / Experimental Thermal and Fluid Science 28 (2004) 407–421

    1o o 1o oU o oU ðrqvUÞK þ ðquUÞK ¼ rC C þ þ SU;K r or oz r or or K oz oz K rU rf SU;K ¼ source terms

M.T. Dhotre et al. / Experimental Thermal and Fluid Science 28 (2004) 407–421

(ii) along the wall: the velocities satisfy the no-slip boundary conditions (the wall function method based on the log law of the wall is invoked to calculate the wall shear stress and the values of k and e close to the wall); (iii) at the inlet: gradients of vG , vL , k and e are set to zero and (iv) at the top surface of the computational domain, the gradient of the dependent variable are set to zero. For initiating the numerical solution, G and uG are specified at the inlet. At all the other locations uG , uL , vG , vL and G are taken to be zero at t ¼ 0. For k and e, the initial guess values were found to be important. 3. Interphase force term The interfacial forces arise due to momentum transfer across the interface. If the slip velocity is constant, the force is called as drag force. If the relative motion is unsteady, virtual mass force prevails in additive to the drag force. When the liquid phase flow pattern is nonuniform in the radial direction the rising bubble experiences a radial (or lateral) force. The formulation of these forces has been discussed in detail by Joshi [5]. In the present work, all the three forces i.e. drag, lift and virtual mass force have been incorporated. 4. Energy balance All the predicted flow patterns must satisfy the energy balance. The rate of energy supply from the gas phase to the liquid phase occurs by two different mechanisms. The rate of energy transfer from mean to turbulent flow is given by: p E ¼ D2 ðqL  qG ÞgHD L ½VG  G VS  ð2Þ 4 The bubble generated turbulent energy, which takes part in momentum transfer, is given by: p ES ¼ CB D2 ðqL  qG ÞgHD L G VS ð3Þ 4 where CB is the fraction of the bubble generated turbulent energy which takes part in the momentum transfer a when CB ¼ 0, the turbulence generated by gas bubbles does not take part in the momentum transfer of the liquid phase. In contrast, when CB ¼ 1, the entire bubble generated turbulence completely participates in the momentum transfer. The rate of total energy transfer is given by the addition of Eqs. (2) and (3) and the result is: p E þ ES ¼ D2 ðqL  qG ÞgHD L ½VG þ ðCB  1ÞG VS  ð4Þ 4 It must be emphasized at this stage that, whatever may be the value of CB , the energy balance must be satisfied. When the k–e model is used for the prediction

411

of flow pattern, we get radial and axial variation of e (energy dissipation rate per unit mass) as one of the answers. From this e field the total energy dissipation rate can be calculated by suitable volume integration. The total energy dissipation rate must equal the energy input rate given by Eq. (4). 5. Drift-flux model The flow pattern in bubble columns mainly depends upon superficial gas velocity, sparger design, liquid and gas phase physical properties. It is possible to represent the combined effect of VG , D, HD , etc., through two parameters: radial hold-up profile and the average bubble rise velocity (VS ). Any change in VG , D, HD and physical properties affect the G profile and VS . Such a relationship is usually expressed through the drift flux model of Zuber and Findlay [6] and is given by the following equation: VG =G ¼ C0 ðVG þ VL Þ þ C1

ð5Þ

where C0 ¼

hG ðG uG þ VL Þi hG ihVG þ VL i

ð6Þ

hG L VS i hG i

ð7Þ

and C1 ¼

The above drift flux model of Zuber and Findlay [6] does not consider the liquid flow pattern within the

Fig. 1. Schematic representation of the development of hold-up profile for (a) single point spargers and (b) multipoint spargers.

412

M.T. Dhotre et al. / Experimental Thermal and Fluid Science 28 (2004) 407–421 0.8 5

1. VG = 0.019

4

2. VG = 0.038 3. VG = 0.064

0.3 3

0.2

AXIAL LIQUID VELOCITY (m/s)

FRACTIONAL GAS HOLD-UP

0.4

4. VG = 0.095 5. VG = 0.169

2 1

0.1

1. VG = 0.019

0.6

4 3 2 1

0.4 0.2 0 0

0.2

0.4

0.6

0.8

-0.4

0.4 2 1

0.2

0.6

0.2

0.4

0.6

0.8

AXIAL LIQUID VELOCITY (m/s)

FRACTIONAL GAS HOLD-UP

3. VG = 0.048 4. VG = 0.096

3

0.1

2 1

0 -0.2

0

0.2

0.4

0.6

0.8

0.2

0.6

0.8

1

Nottenkamper et al. (1983)

DIMENSIONLESS RADIAL DISTANCE

0.8

4 3

1. VG = 0.012 2. VG = 0.024 3. VG = 0.048

0.6 0.4 0.2 0 -0.2 0

2

4. VG = 0.096

1

0.2

0.4

0.6

0.8

1

-0.4 -0.6 Menzel et al. (1990)

-1

1

(c)

DIMENSIONLESS RADIAL DISTANCE

0.4

-0.4

-0.8

0 0

2. VG = 0.105 3. VG = 0.324

0.2

1 0.2

1. VG = 0.053

2 1

(b)

1. VG = 0.012 2. VG = 0.024

1

-0.8

1

DIMENSIONLESS RADIAL DISTANCE

4

3

0.4

-0.6 0 0

0.8

DIMENSIONLESS RADIAL DISTANCE

0.8 AXIAL LIQUID VELOCITY (m/s)

FRACTIONAL GAS HOLD-UP

0.6

3

0.6

Hills (1974)

(a)

1. VG = 0.053 2. VG = 0.105 3. VG = 0.324

0.4

-0.6

1

DIMENSIONLESS RADIAL DISTANCE

0.2

-0.2

0 0

2. VG = 0.038 3. VG = 0.064 4. VG = 0.095 5. VG = 0.169

5

DIMENSIONLESS RADIAL DISTANCE

Fig. 2. Comparison between the simulated and experimental profiles of fractional gas hold up and axial liquid velocity (refer to Table 3 for experimental details). (a) Hills [12]; (b) Nottenkamper et al. [13]; (c) Menzel et al. [14]; (d) Yao et al. [15]; (e) Yu and Kim [16]; (f) Grienberger and Hofmann [17].

column. As a result, for practically all the systems, the value of C0 estimated by Eq. (6) is always much lower than that obtained from a plot of VG =G versus (VG þ VL ) according to Eq. (5). Therefore, Ranade and Joshi [7] and Joshi et al. [8] modified the drift flux model by including the radial profile of liquid velocity. The definition of C1 remains the same as Eq. (7). The parameter C0 was modified to: C0 ¼

hG ðG uG þ VL Þi hG L uL i þ hG ihVG þ VL i hG ihVG þ VL i

ð8Þ

The parameters C0 and C1 are the drift flux constants. C0 represents the hold-up profile and C1 the bubble rise velocity. The most fortunate characteristic feature of bubble columns is that the values of C0 and C1 are practically independent of the column diameter (of course when D > 150 mm and the sparger region is exceeded). Therefore, for a given gas–liquid system, a few measurements of G with respect to VG and VL (over the range of interest) in a small diameter column (150 mm) enable the estimation of C0 and C1 .

M.T. Dhotre et al. / Experimental Thermal and Fluid Science 28 (2004) 407–421

4

AXIAL LIQUID VELOCITY (m/s)

0.6

0.3 FRACTIONAL GAS HOLD-UP

413

1. VG = 0.02 2. VG = 0.04

3

3. VG = 0.06 4. VG = 0.08

0.2 2

1

0.1

0.4

4 3 2

0.2

1

1. VG = 0.02 2. VG = 0.04 3. VG = 0.06 4. VG = 0.08

0 0

0.2

0.4

0.6

0.8

1

-0.2 -0.4 Yao et al. (1991)

0

0

0.2

0.4

0.6

0.8

1

(d)

DIMENSIONLESS RADIAL DISTANCE

0.12

AXIAL LIQUID VELOCITY (m/s)

1. V G = 0.01 2. V G = 0.02 3. V G = 0.04 4. V G = 0.06 5. V G = 0.10

0.08

0.04

0 0

0.2 0.4 0.6 0.8 DIMENSIONLESS RADIAL DISTANCE

AXIAL LIQUID VELOCITY (m/s)

0.25 0.2 0.15

1

0.1

0.2 0.1 0 -0.1

0

0.2

0.05

0.4

0.6

0.8

1

-0.2 -0.3

(e)

1. VG = 0.02 2. VG = 0.08

2

1. V G = 0.01 2. V G = 0.02 3. V G = 0.04 4. V G = 0.06 5. V G = 0.10

0.3

-0.4

1

0.3 FRACTIONAL GAS HOLD-UP

DIMENSIONLESS RADIAL DISTANCE 0.4

0.16 FRACTIONAL GAS HOLD-UP

-0.6

DIMENSIONLESS RADIAL DISTANCE 0.5 2. VG = 0.08

0.3 1

0.2 0.1 0

-0.1

0

0.2

0.4

0.6

0.8

1

-0.2 -0.3

0

1. VG = 0.02

2

0.4

Grienberger and Hofmann (1992)

-0.4 0

0.2

0.4

0.6

0.8

1

(f)

DIMENSIONLESS RADIAL DISTANCE

DIMENSIONLESS RADIAL DISTANCE

Fig. 2 (continued)

The flow in the 0.385 m diameter cylindrical bubble column has been simulated. The simulations are compared with the experiments performed by Parasu Veera and Joshi [1,3]. They have studied the effect of the sparger design, height of dispersion and liquid phase properties on the fractional gas hold-up. The effect of sparger design on the development of radial hold-up profiles is shown schematically in Fig. 1. 6. Experimental apparatus Parasu Veera and Joshi [1,3] performed experiments in a Perspex cylindrical bubble column of 0.385 m i.d.

and 3.2 m height. Five different sieve plate spargers were employed with free area in the range of 0.138–5.4% and hole diameter in the range of 0.8–87 mm. Experiments were carried out for five different gas velocities and air flow rate was measured with a pre-calibrated rotameter. Experiments were carried out at three axial locations, one at just above the sparger HD =D ¼ 0:259 and other at HD =D ¼ 3 and 5. The height of the dispersion was maintained constant (HD =D ¼ 6). The average fractional gas hold-up in all cases was estimated from knowledge of the bed expansion, while a gamma ray tomography system was used for the measurement of radial hold-up profiles.

414

M.T. Dhotre et al. / Experimental Thermal and Fluid Science 28 (2004) 407–421

Table 2 Comparison of radial gas hold-up and mean axial liquid velocity profiles: Experimental data of previous work Superficial gas velocity (mm/s)

Measurement location HD =D

Measurement techniques employed Hold-up

Liquid velocity

Air–water

19–169

4.34

Pavlov tube

Air–water

53–1452



Electro-conductivity needle probe Optical probe

Air–water

12–96

4.86

Air–deionized water

20–80

3.49

Perforated pipes: Air–water 6 mm, 8 nos. hole diameter––1 mm, 78 nos.

10–140

0.475

4.50

Hole diameter–– 0.2 mm

Air–water

20–80

3.5

Height measurement



Single point (d0 ¼ 25 mm), Multipoint (d0 ¼ 3 mm, 71 nos.)

Air–water Air–alcohol Air–CMC

60–290

0.259 3.0 5.0

Gamma ray tomography

Column diameter (m)

Column height (m)

Sparger

Hills [12]

0.138

1.37

Nottenkamper et. al. [13] Menzel et al. [14]

0.45



0.6

3.45

Seive plate: Hole diameter––0.4 m, 61 nos. Seive plate hole diameter––1 mm –

Yao et al. [15]

0.290

4.50

Hole diameter–– 0.2 mm

Yu and Kim [16]

0.254

2.5

Grienberger and Hofmann [17]

0.290

Parsu Veera and Joshi [1,3]

0.385

Researchers

Gas–liquid system

7. Numerical modeling 7.1. Solution procedure The set of steady state governing equations given in Table 1 were solved numerically and involved the following steps: (i) generation of suitable grid system, (ii) conversion of governing equation into algebraic equations, (iii) selection of discretization schemes, (iv) formulation of the discretized equation at every grid location, (v) formulation of pressure equation, (vi) development of a suitable iteration scheme for obtaining a final solution. The finite control volume technique of Patankar [9] was employed for the solution of these equations. A staggered grid arrangement proposed by the Patankar and Spalding [10], consisted of 64 · 100 grid points, with 64 grid points in radial direction and 100 grid point in axial direction. The velocity components were calculated for the points that lie on the faces of the control volume, while all other variables were calculated at the center of the control volume. A power law scheme was used for the discretization of the governing equations. A SIMPLE algorithm was used to solve the pressure velocity coupling term. The set of algebraic equations obtained

Electro-conductivity microprobe Electro-conductivity microprobe Two channel optical fibre probe

Fly-wheel anemometer Hot film anemometry Hot film anemometry Two Pt–Rh electrodes, two U shaped optical fibres and a tracer injection Five point conductivity probe –

after discretization were solved by TDMA. Relaxation parameters and internal iterations for the variables were tuned to optimize the balance between the convergence criteria (1.0 · 103 ) and the number of iterations required. The detailed stepwise procedure for getting the flow pattern is given by Joshi [5]. 7.2. Modeling of sparger region The first step to model the sparger region is to resolve the near sparger volume with very fine grids. The second step is to specify the orifice velocity of the gas which can be calculated from the percentage free area of the sparger as: VO ¼ VG =%FA

ð9Þ

The resistance offered by the sparger results in a pressure drop, and Thorat et al. [11] have given a correlation for the evaluating pressure drop across the sparger, i.e. DpW ¼ aVO ðrÞ þ bVO2 ðrÞ

ð10Þ

The coefficients a and b are dimensional proportionality constants that can be determined either

M.T. Dhotre et al. / Experimental Thermal and Fluid Science 28 (2004) 407–421

415

Table 3 Comparison of CFD predictions at HD =D ¼ 5 with experimental observations Authors

VG (m/s)

G

Hills [12]

0.0206 (0.019) 0.038 (0.038) 0.067 (0.064) 0.106 (0.095) 0.171 (0.169)

0.065 0.106 0.161 0.197 0.218

Nottenkamper et al. [13]

0.054 (0.053) 0.106 (0.105) 0.331 (0.324)

Menzel et al. [14]

0.012 0.026 0.052 0.096

Yao et al. [15]

0.021 0.041 0.062 0.081

Energy balancea

Material balance C1

LHS

RHS

(2.0) (2.0) (2.0) (2.0) (2.0)

0.222 (0.25) 0.24 (0.25) 0.25 (0.25) 0.32 (0.25) 0.465 (0.25)

0.105 0.127 0.210 0.355 0.857

0.115 0.205 0.221 0.366 0.831

0.121 (0.120) 0.175 (0.168) 0.312 (0.271)

2.64 (2.0) 2.5 (2.0) 1.64 (2.0)

0.375 (0.40) 0.35 (0.40) 0.54 (0.40)

0.190 0.502 1.815

0.198 0.513 1.836

(0.012) (0.024) (0.048) (0.096)

0.029 0.049 0.105 0.137

(0.029) (0.048) (0.103) (0.128)

3.13 5.80 5.30 3.02

(2.5) (2.5) (2.5) (2.5)

0.39 0.36 0.42 0.35

0.057 0.130 0.223 0.593

0.07 0.157 0.263 0.588

(0.02) (0.04) (0.06) (0.08)

0.057 0.108 0.156 0.186

(0.057) (0.106) (0.153) (0.180)

2.96 2.71 2.28 2.69

(2.0) (2.0) (2.0) (2.0)

0.285 0.236 0.251 0.223

(0.25) (0.25) (0.25) (0.25)

0.075 0.208 0.299 0.417

0.078 0.233 0.312 0.447

Yu and Kim [16]

0.011 (0.01) 0.02 (0.02) 0.040 (0.04) 0.061 (0.06) 0.101 (0.1) 0.142 (0.14)

0.025 0.035 0.049 0.062 0.090 0.136

(0.025) (0.034) (0.048) (0.060) (0.084) (0.130)

3.17 (2.1) 2.4 (2.1) 2.5 (2.1) 2.23 (2.1) 1.87 (2.1) 1.77 (2.1)

0.39 (0.40) 0.554 (0.73) 0.739 (0.73) 0.853 (0.73) 0.76 (0.73) 0.80 (0.73)

0.045 0.122 0.187 0.322 0.526 0.661

0.054 0.100 0.194 0.288 0.549 0.668

Grienberger and Hofmann [17]

0.021 (0.02)

0.057 (0.054)

2.96 (2.0)

0.285 (0.25)

0.075

0.078

0.081 (0.08)

0.156 (0.150)

2.70 (2.0)

0.22 (0.25)

0.417

0.447

C0

a

(0.067) (0.107) (0.160) (0.194) (0.212)

3.84 3.34 2.58 2.10 1.91

(0.40) (0.40) (0.40) (0.40)

The bracketed quantities are the experimental values; LHS is volume integral of e and RHS is Eq. (4).

Table 4 Comparison of CFD predictions with experimental observations for multipoint sparger (do ¼ 3 mm, FA ¼ 0.42%) System

VG

Air–water

0.06 0.12 0.18 0.24 0.29

Air–alcohol

Air–CMC

a

Energy balancea

Material balance

G

CB

LHS

RHS

C0

C1

(0.06) (0.12) (0.18) (0.24) (0.29)

0.13 (0.13) 0.191(0.19) 0.241 (0.24) 0.268 (0.268) 0.285 (0.286)

0.50 0.45 0.21 0.15 0.42

0.353 0.533 0.911 1.352 1.50

0.345 0.521 0.899 1.325 1.452

2.36 (2.36) 2.4 (2.36) 2.2 (2.36) 2.35 (2.36) 2.3 (2.36)

0.31 0.34 0.35 0.33 0.33

0.06 0.12 0.18 0.24 0.29

(0.06) (0.12) (0.18) (0.24) (0.29)

0.192 0.245 0.282 0.302 0.358

(0.1987) (0.249) (0.285) (0.318) (0.364)

0.15 0.20 0.14 0.40 0.50

0.32 0.55 0.85 1.45 1.72

0.312 0.541 0.75 1.422 1.658

1.8 (1.96) 1.92 (1.96) 2.0 (1.96) 1.98 (1.96) 1.96 (1.96)

0.2 (0.22) 0.25 (0.22) 0.28 (0.22) 0.28 (0.22) 0.23 (0.22)

0.06 0.12 0.18 0.24 0.29

(0.06) (0.12) (0.18) (0.24) (0.29)

0.089 0.128 0.180 0.195 0.235

(0.09) (0.139) (0.181) (0.20) (0.239)

0.15 0.21 0.30 0.14 0.23

0.29 0.51 0.73 1.12 1.82

0.315 0.582 0.765 1.215 1.921

2.50 2.56 2.45 2.30 2.30

0.51 0.56 0.58 0.58 0.54

(2.48) (2.48) (2.48) (2.48) (2.48)

(0.33) (0.33) (0.33) (0.33) (0.33)

(0.55) (0.55) (0.55) (0.55) (0.55)

The bracketed quantities are the experimental values; LHS is volume integral of e and RHS is Eq. (4).

experimentally or by using empirical relations. As outlined earlier two-sparger plates have been used in the present work. The values of a and b were estimated

using the procedure given by Thorat et al. [11]. ((i) Multipoint sparger: a ¼ 32:12, b ¼ 0:404; (ii) Single point sparger a ¼ 24:32, b ¼ 0:512).

416

M.T. Dhotre et al. / Experimental Thermal and Fluid Science 28 (2004) 407–421

Table 5 Comparison of CFD predictions with experimental observations for single point sparger (do ¼ 25 mm, FA ¼ 0.42%) Energy balancea

Material balance

LHS

RHS

C0

C1

0.312 0.497 0.879 0.982

0.304 0.418 0.862 0.925

2.8 (2.5) 2.65 (2.5) 2.4 (2.5) 2.6 (2.5)

0.32 (0.368) 0.358 (0.368) 0.354 (0.368) 0.320 (0.368)

0.10 0.15 0.20 0.15

0.383 0.512 0.921 1.212

0.42 0.59 0.89 0.968

2.1 2.3 2.5 2.2

(2.3) (2.3) (2.3) (2.3)

0.251 (0.252) 0.251 (0.252) 0.27 (0.252) 0.281 (0.252)

0.21 0.22 0.12 0.25

0.311 0.592 0.896 1.412

0.351 0.561 0.861 1.345

3.3 3.8 3.6 3.2

(3.6) (3.6) (3.6) (3.6)

0.54 0.62 0.61 0.59

System

VG

Air–water

0.06 0.12 0.18 0.24

(0.06) (0.12) (0.18) (0.24)

0.122 0.177 0.224 0.254

(0.12) (0.175) (0.22) (0.251)

0.27 0.15 0.18 0.14

Air–alcohol

0.06 0.12 0.18 0.24

(0.06) (0.12) (0.18) (0.24)

0.159 0.227 0.250 0.291

(0.159) (0.228) (0.252) (0.309)

Air–CMC

0.06 0.12 0.18 0.24

(0.06) (0.12) (0.18) (0.24)

0.080 0.115 0.138 0.174

(0.082) (0.111) (0.135) (0.175)

a

CB

G

(0.58) (0.58) (0.58) (0.58)

The bracketed quantities are the experimental values; LHS is volume integral of e and RHS is Eq. (4).

FRACTIONAL GAS HOLD-UP

0.4

1

update the orifice velocity. For this purpose, the following relation has been used:

1. Air-Alcohol 2. Air-Water 3. Air CMC

2 0.3

pU  pD ðrÞ ¼ aVO ðrÞ þ bVO2 ðrÞ

3

where, pD ðrÞ is the pressure at the bottom of the column (down stream the sparger) at any radial location r. As an initial guess, the hold-up at the holes is uniformly specified. For the case of single point sparger, few cells in the center of the columns were made Ôlive’ and the orifice velocity was specified in those cells. The gas holdup in the live cells was specified as unity and zero in the Ôdead’ cells.

0.2

0.1

0 0

AXIAL LIQUID VELOCITY (m/s)

0.7 0.5

0.2 0.4 0.6 0.8 DIMENSIONLESS RADIAL DISTANCE

3 2

1

1. Air-Alcohol 2. Air-Water 3. Air-CMC

1

ð12Þ

8. Results and discussion 8.1. Comparison of the flow pattern with experimental data

0.3 0.1 -0.1 0

0.2

0.4

0.6

0.8

1

-0.3 -0.5 DIMENSIONLESS RADIAL DISTANCE

Fig. 3. Effect of liquid phase physical properties on the fractional gas hold-up and liquid phase velocity profile for superficial velocity of 0.12 m/s (also refer Table 5).

The upstream pressure (below the sparger plate) can be calculated as pU ¼ ðL qL þ G qG ÞgHD þ DpW

ð11Þ

The calculated orifice velocity was used as a boundary condition at the orifice location. The downstream pressure profile obtained from CFD simulations is used to

As a first step, it is important to establish the validity of the model for flow pattern. Comparison has been made with the experimental data of Hills [12], Nottenkamper et al. [13], Menzel et al. [14], Yao et al. [15], Yu and Kim [16], Grienberger and Hoffman [17] and shown in the Fig. 2(a–f) and in Table 3. The experimental details of these cases are given in Table 2. The agreement between the predicted and the experimental profiles can be seen to be excellent. The agreement over the wide range of D, VG and the gas–liquid systems ensures the applicability of the model for the estimation of flow pattern. 8.2. Effect of physical properties The effect of the physical properties of the liquid phase was studied. As can be observed from the Tables 4

M.T. Dhotre et al. / Experimental Thermal and Fluid Science 28 (2004) 407–421

VG=0.12 m/s

VG=0.06 m/s

0.4

0.3

0.3

0.2

0.2

0.1

0.1 HD /D=5

0.2 0.1 H D /D=3

0

HD /D=5

0

HOLD-UP

0

HOLD-UP

417

0.4 0.3 0.2 0.1

HD /D=3

0

0.4 0.2

0.3 0.2

0.1

0.1

H D /D=0.259

0 -1

(a)

HD /D=0.259

0 -0.5

0

0.5

-1

1

DIMENSIONLESS RADIAL DISTANCE

(b)

-0.5

0

0.5

1

DIMENSIONLESS RADIAL DISTANCE

VG=0.18 m/s

0.4 0.3 0.2 0.1

HD /D=5

HOLD-UP

0 0.4 0.3 0.2 0.1

HD /D=3

0

0.4 0.3 0.2 HD /D=0.259

0.1 0 -1

(c)

-0.5

0

0.5

1

DIMENSIONLESS RADIAL DISTANCE

Fig. 4. Radial gas hold-up profiles at various axial locations for various liquid phases for multipoint sparger plate. (a) VG ¼ 0:06 m/s, (b) VG ¼ 0:12 m/s, (c) VG ¼ 0:18 m/s, (d) VG ¼ 0:24 m/s, (e) VG ¼ 0:29 m/s; (N) air–alcohol, (d) air–water, () air–CMC.

and 5, the average hold-up increased for the air–alcohol system while it decreased for the CMC solution. Fig. 3 compares the effect of different gas–liquid systems on the flow pattern for the superficial velocity of 0.12 m/s. It is known that the presence of alcohol reduces the average bubble size and the rise velocity as compared to that for the air–water system. In contrast, the non-Newtonian nature of the liquid phase increases the occurrence of coalescence and the values of dB and slip velocities are higher than those for the air–water system (Haque et al.

[18]). The experimental values of G , C0 and C1 are given in Table 4 and 5. It can be seen that, in the presence of an alcohol, the value of slip velocity (C1 ¼ 0:22) is lower and the fractional gas hold-up is higher than the corresponding values for the air–water system (C1 ¼ 0:33). Further, the hold-up profile is flatter (C0 ¼ 1:96) than that in the air–water system (C0 ¼ 2:36). Therefore, the liquid circulation velocities are lower in the presence of an alcohol solution. It can be seen from Tables 4 and 5, that, in the case of aqueous solution of CMC, the value

418

M.T. Dhotre et al. / Experimental Thermal and Fluid Science 28 (2004) 407–421

VG=0.24 m/s

0.5

VG=0.29 m/s

0.5

0.4

0.4

0.3

0.3

0.2

0.2

0.1

0

0.4

HOLD-UP

HOLD-UP

0

0.3 0.2 0.1

H D /D=5

0.1

HD /D=5

HD /D=3

0

0.5 0.4 0.3 0.2 0.1 0

H D /D=3

0.5 0.4

0.4

0.3

0.3

0.2 0.1 0 -1

(d)

0.2 H D /D=0.259

0.1

HD /D=0.259

0 -0.5 0 0.5 1 DIMENSIONLESS RADIAL DISTANCE

-1

(e)

-0.5 0 0.5 1 DIMENSIONLESS RADIAL DISTANCE

Fig. 4 (continued)

of the slip velocity (C1 ¼ 0:55) is higher and the fractional gas hold-up is lower than those for the air–water system. Further, the hold-up profile is steeper (C0 ¼ 2:48) as compared to that of the air–water system (C0 ¼ 2:36). This means that the driving force for the liquid circulation is higher and results in a relatively more intense liquid circulation.

circulating bubbles. This could not be predicted in the CFD predictions, which are given in the Table 5. In the center region up to r=R range of 0.8, excellent agreement can be seen from the figures.

8.3. Radial hold-up profiles at the axial location of H =D ¼ 0:259

While describing the hold-up profiles, we have used two terms: flat and steep. Flat profiles mean practically no variation in hold-up in the radial direction. On contrast, steep profiles mean the existence of high centerline hold-up and steep reduction as the wall is radially approached. For multipoint spargers, the profiles at this axial location were relatively steeper than at the axial location of 0.259 (Fig. 4(a–e)) for air water and aqueous CMC solution. Whereas for the aqueous alcohol solution, the profiles are practically similar to the profiles at the axial location of 0.259. For the single point sparger, the profiles were found to become flatter with an increase in distance from the sparger (Fig. 5(a–d)). The change in the centerline hold-up for the single point sparger was significant. Good agreement between the experimental and predicted values of the hold-up profiles was obtained.

Figs. 4(a–e) and 5(a–d) show that the gas hold-up profiles at the axial locations for the water, aqueous solutions of alcohol and CMC for multipoint and single point spargers, respectively. For the multipoint sparger, it can be seen that three distinct profiles are obtained for the three liquid phases, details are given in Table 4. The profiles appear to be relatively flat. At any radial location, the air–alcohol system exhibits the highest value of G and the air–aqueous CMC system exhibits the lowest. The G profile for the air–water system lies in between. This is due to the bubble sizes generated at the sparger. In the presence of small quantities of alcohol, smaller bubbles are generated due to the surface active property of the alcohols. In contrast, the aqueous solution of CMC has high viscosity and relatively large bubbles are generated as compared to those in water. For the single point sparger (Fig. 5(a–d)), the profiles for all three liquid systems are fairly close to each other. The profiles are very steep at the centre and become practically zero in the r=R range of 0.7–1. However, near the wall (0:9 < r=R < 1:0), G was found to increase in the experimental data, which is probably due to the re-

8.4. Radial hold-up profiles at the axial location of HD =D ¼ 3

8.5. Radial hold-up profiles at axial location of HD =D ¼ 5 At this axial location for the aqueous solution of alcohol, the profiles become steep compared to very flat for the other two axial locations (0.259 and 3) for the multipoint sparger (Fig. 4(a–e)). Similarly for the

M.T. Dhotre et al. / Experimental Thermal and Fluid Science 28 (2004) 407–421

0.3

VG=0.12 m/s

0.4

VG=0.06m/s

419

0.3

0.2

0.2 0.1

0.1

HD /D=5

0.4 HOLD UP

0.3 HOLD-UP

HD /D=5

0

0

0.2 0.1

0.3 0.2 0.1

HD /D=3

0

0

0.3

0.4

0.2

0.3

HD /D=3

0.2 0.1

HD /D=0.259

0.1

HD /D=0.259

0

0 -1

(a)

-0.5

0

0.5

-1

1

DIMENSIONLESS RADIAL DISTANCE

-0.5

(b)

0

1

VG=0.24 m/s

VG=0.18 m/s

0.4

0.4

0.3

0.2

0.2 0.1

0.5

DIMENSIONLESS RADIAL DISTANCE

H D /D=5

0

H D /D=5

0.5 0.4 0.3 0.2 0.1 0

HOLD-UP

HOLD-UP

0

H D /D=3

0.5

0.5

0.4

0.4

0.3

0.3

HD /D=0.259

0.1

H D /D=3

0.2 0.1

0.2

H D /D=0.259

0

0

(c)

0.5 0.4 0.3 0.2 0.1 0

-1

-0.5 0 0.5 1 DIMENSIONLESS RADIAL DISTANCE

-1

(d)

-0.5

0

0.5

1

DIMENSIONLESS RADIAL DISTANCE

Fig. 5. Radial gas hold-up profiles at various axial locations for various liquid phases for single point sparger plate. (a) VG ¼ 0:06 m/s, (b) VG ¼ 0:12 m/s, (c) VG ¼ 0:18 m/s, (d) VG ¼ 0:24 m/s; (N) air–alcohol, (d) air–water, () air–CMC.

aqueous solution of CMC a similar profile was obtained at this axial location. But the variation with axial location of HD =D ¼ 3 is very nominal. In the case of the single point sparger, the profile flattens at this axial location in both the liquid phases Fig. 5(a–d). It is worth noting that for a single point sparger, for an aqueous solution of alcohol, the profiles are practically the same at both the axial locations HD =D ¼ 3 and 5, and for a multipoint sparger at both the axial locations HD =D ¼ 0:259 and HD =D ¼ 3. For the air–CMC system, the profiles vary from HD =D ¼ 0:259 to HD =D ¼ 3 considerably. This clearly shows that the effect of the

sparger is very dominating until HD =D ¼ 3 in the case of the air–aqueous alcohol system, whereas in the air– aqueous CMC system, the sparger effect is vanishing at this aspect ratio. The same trend was observed in the CFD predictions, as shown in Figs. 4(a–e) and 5(a–d). 8.6. Development of hold-up profiles in the column The development of the hold-up profile for both multipoint and single point spargers at superficial gas velocities of 0.06, 0.12, 0.18, 0.24, 0.29 m/s are shown in Figs. 4(a–e) and 5(a–d) for the air–water, air–aqueous

420

M.T. Dhotre et al. / Experimental Thermal and Fluid Science 28 (2004) 407–421

solution of alcohol and air-aqueous solution of CMC system. It can be seen from Fig. 4(a–e), for multipoint spargers, the profile is relatively flat at the HD =D ratio of 0.259 and increases towards the centerline as height increases from sparger. Hence the driving force for the liquid circulation, DG , (the difference between the centerline hold-up and wall hold-up) increases as the distance increases from the sparger. Therefore, the average hold-up decreases as the dispersion height to diameter ratio (HD =D) increases for multipoint spargers. The opposite trend was observed for the single point sparger (Fig. 5(a–d)), in that the profile is steep near the sparger and it becomes flat as the distance increases from the sparger, thereby decreasing the driving force for liquid circulation. Hence, the average hold-up increases as the dispersion height to diameter ratio increases for single point spargers. Similar trends were observed for aqueous solutions of alcohol and CMC for single and multipoint spargers, as shown in Figs. 4(a–e) and 5(a–d).

9. Conclusions 1. A stepwise procedure has been developed for the prediction of radial profiles of gas hold-up and axial liquid phase velocity. An excellent agreement between predicted and experimental profiles of hold-up and velocity was observed for a wide range of column diameter 0:138 < D < 0:6 m, column height 1:37 < HD < 4:5 m, and superficial gas velocity (0:01 < VG < 0:324, m/s). These data have been reported in different laboratories by Hills [12], Nottenkamper et al. [13], Menzel et al. [14], Yao et al. [15], Yu and Kim [16], Grienberger and Hofmann [17]. The model was also successfully applied to system involving air–water, air–aqueous solutions of alcohol and air– CMC solutions. 2. The predicted gas hold-up profiles by CFD simulation were found to agree with experimental measurements in the 0.385 m i.d. column. Comparisons were made at three locations, two-sparger designs, three gas–liquid systems and a wide range of superficial gas velocity (0.06–0.29 m/s). 3. For multipoint spargers, similar to the air–water system, the centreline hold-up increases and the wall hold-up decreases with an increase in distance from the bottom. For multipoint spargers (FA ¼ 0.42%), the profiles were relatively flat at the bottom and increased towards the centerline they become steeper with an increase in the height of dispersion. 4. For the single point sparger of 25 mm hole diameter, the gas hold-up profiles were very steep at the axial location of HD =D ¼ 0:259. They then became flatter with an increase in distance from the sparger. The profiles were found to be independent of the liquid

phase properties at this location for the single point sparger. 5. With an increase in height of dispersion, for multipoint spargers, the profiles undergo a shape change of flatter to steeper near the sparger. However, this type of variation occurs up to an HD =D ratio of 5 for the air–water, and up to HD =D ¼ 3 for air– CMC solutions system. Beyond these values, the gas hold-up profiles were practically the same irrespective of the sparger design, and remained relatively unchanged with further increase in HD =D. For air–aqueous solution of alcohol system, the profile remained flat even beyond HD =D ¼ 5, suggesting that the sparger influence remains up to a much higher HD =D ratio, as compared with air–water and air– aqueous solution of CMC systems. References [1] U. Parasu Veera, J.B. Joshi, Measurement of gas hold up profiles by gamma ray tomography: Effect of sparger design and height of dispersion in bubble columns, Transactions of Institution of Chemical Engineers 77 (1999) 303–317. [2] J. Zahradnik, M. Fialova, F. Kartanck, K.D. Green, N.H. Thomas, The effect of electrolytes on bubble coalescence and gas hold-up in bubble column reactors, Transactions of Institution of Chemical Engineers 73 (1995) 341–346. [3] U. Parasu Veera, J.B. Joshi, Measurement of gas hold up profiles in bubble column by gamma ray tomography: Effect of Liquid phase properties, Transactions of Institution of Chemical Engineers 78 (2000) 303–317. [4] J.B. Joshi, U. Parasu Veera, CH.V. Prasad, D.V. Phanikumar, N.S. Deshpande, S.S. Thakre, B.N. Thorat, Gas hold-up structures in bubble column reactors, PINSA 64 (1998) 441–567. [5] J.B. Joshi, Computational flow modelling and design of bubble column reactors, Chemical Engineering Science 56 (2001) 5893– 5933. [6] N. Zuber, J.A. Findlay, Average volumetric concentration in two phase flow systems, Journal of Heat Transfer 87 (1969) 453–468. [7] V.V. Ranade, J.B. Joshi, Transport phenomena in multiphase reactors, in: Proceedings Inst Symposium on Transport Phenomena in Multiphase Systems, BHU Press, Varanasi, 1987, pp. 113–196. [8] J.B. Joshi, V.V. Ranade, S.D. Gharat, S.S. Lele, Sparged loop reactors, Candian Journal of Chemical Engineering 68 (1990) 705– 741. [9] S.V. Patankar, Numerical Heat Transfer and Fluid Flow, McGraw-Hill, New York, 1980. [10] S.V. Patankar, D.B. Salding, A calculation procedure for heat, mass and momentum transfer in three dimensional parabolic flows, International Journal of Heat and Mass Transfer 15 (1972) 126–129. [11] B.N. Thorat, K. Kataria, A.V. Kulkarni, J.B. Joshi, Pressure drop studies in bubble columns, Industrial Engineering Chemistry and Research 40 (2001) 3675–3688. [12] J.H. Hills, Radial non uniformity of velocity and voidage in a bubble column, Transactions of Institution of Chemical Engineers 52 (1974) 1–9. [13] R. Nottenkamper, A. Stieff, P.M. Weinspach, Experimental investigations of hydrodynamics of bubble columns, German Chemical Engineering 6 (1983) 147–155. [14] T. Menzel, T. Weide, O. Staudacher, O. Wein, U. Onken, Reynolds shear stress for modeling of bubble column reactor, Industrial Engineering Chemistry and Research 29 (1990) 988–994.

M.T. Dhotre et al. / Experimental Thermal and Fluid Science 28 (2004) 407–421 [15] B.P. Yao, C. Zheng, H.E. Gasche, H. Hofmann, Bubble behaviour and flow structure of bubble columns, Chemical Engineering Processing 29 (1991) 65–75. [16] Y.H. Yu, S.D. Kim, Bubble properties and local liquid velocity in the radial direction of co-current gas–liquid flow, Chemical Engineering Science 46 (1991) 313–320.

421

[17] J. Grienberger, H. Hofmann, Investigation and modeling of bubble columns, Chemical Engineering Science 47 (1992) 2215–2220. [18] M.W. Haque, K.D.P. Nigam, K. Viswanathan, J.B. Joshi, Studies on bubble rise velocity in bubble columns employing nonNewtonian solutions, Chemical Engineering Communications 73 (1988) 31–42.

CFD simulation of sparger design and height to ...

+91-22-2414-5616; fax: +91-22-2414-. 5614. ... radial virtual mass force per unit volume of .... fore, the drag force (FDZ and FDR), virtual mass force (Fvz and Fvr) ...

467KB Sizes 1 Downloads 227 Views

Recommend Documents

Simulation of oscillatory baffled column: CFD and ...
+1 780 264 7201; fax: +1 780 492 2881. E-mail address: ... spacing and baffle free area) conditions (Ni et al., 1995). The flow passing through the ... With the continuous ad- vancement of computer technologies, the use of CFD methodol-.

Advancing Thermoacoustics Through CFD Simulation ...
alternative that can provide cooling to essentially any required temperature level ..... The Energy Information Agency published information on the number of ...

Advancing Thermoacoustics Through CFD Simulation ...
The 2007 Buildings Energy Data Handbook (DOE) illus- trates that CFC-12 is ... the critical temperature gradient for thermoacoustic engines. The expression for ...

CFD simulation of a thermoacoustic engine with coiled ...
Oct 17, 2009 - CFD simulation of a thermoacoustic engine with coiled resonator☆ .... developed a computational fluid dynamics (CFD) simulation of a stand-.

CFD simulation of a thermoacoustic engine with coiled ...
Oct 17, 2009 - The transition to thermoacoustic technology occurred when Ceperley .... condition on the compliance side was replaced by another adiabatic.

CFD simulation of homogeneous reactions in turbulent ...
+91 22 414 5616; fax: +91 22 414 5614. E-mail address: ..... [5] F. Durst, J. Jovanovic, J. Sender, LDA measurements in the near-wall region of a turbulent pipe ...

cfd simulation of flow and axial dispersion in external ...
ses rapidly at low VG range (up to 0.02 m sJ1) and tend to level off with .... The model constants are Cm ј 0.09; Cm,BI ј 0.6; sk ј 1.00; s1 ј 1.00; C11 ј 1.44, C12 ј ...

Design of a gas distributor: Three-dimensional CFD ...
+91 22 414 5616; fax: +91 22 414 5614. ... axial virtual mass force per unit volume of disper- .... force (FDZ, FDR and FDθ), virtual mass force (FVZ, FVR, FVθ),.

The limits to tree height
2Department of Biological Sciences, Humboldt State University, Arcata,. California ... tension and therefore leaf water potential (W) predicted for great heights ... Humboldt Redwoods State Park, California. a, Xylem pressure of small branches.

The limits to tree height
The regression line is fitted to data from six trees. .... Great height may prevent recovery of lost hydraulic function by embolism ..... Kramer, P. J. & Boyer, J. S. Water Relations of Plants and Soils (Academic, San Diego, 1995). 11. Taiz, L.

CFD Simulation of Heat Transfer in Turbulent Pipe Flow
radial variation of axial velocity, the turbulent kinetic energy, and the eddy diffusivity (compared with near- wall experimental data of Durst et al.5). The fourth.

CFD Simulation of the Flow Pattern for Drag Reducing ...
criteria; accurate prediction of the radial variation of axial velocity, the turbulent kinetic energy and the eddy diffusivity (compared with the experimental data of.

Design, Simulation and Implementation of a MIMO ...
2011 JOT http://sites.google.com/site/journaloftelecommunications/. Design, Simulation and Implementation of a. MIMO MC-CDMA based trans-receiver system.

Design and Simulation of Multi-Band Microstrip Patch Antenna ... - IJRIT
resulting structure is less rugged but has a wider bandwidth. Because ... Figure2.1: Top View of Multi-band Antenna and Measurement Specifications. Primarily ...

Design, Simulation and Testing of an Optimized ...
Design, Simulation and Testing of an Optimized ... literatures and data obtained from various pathological tests, ... (nonlinear) mapping from input to output.

Design, Simulation and Performance of Reflecting ...
Jul 18, 2006 - Concentrator or optical system is the part of the collector that directs radiation onto the receiver. • Area concentration ratio C or geometric ...

Design and Simulation of an Easy Structure Multiband ...
Abstract— An easy structure printed slot antenna providing multi frequency is designed and simulated. ... line. To match the input impedance of the antenna to the 50 ohm SMA ... monopole slot antennas for mobile phone applications.

Design and Simulation of Adaptive Filtering Algorithms for Active ...
Keywords: Adaptive Filter, LMS Algorithm, Active Noise cancellation, MSE, .... The anti-noise generated corresponding to the white noise is as shown below,.

Design and Simulation of Adaptive Filtering Algorithms ...
masking speech than broadband noise, the degree of masking varies with ... Also passive methods work quite well only for frequencies above 500 Hz and active ...

Design and Simulation of Adaptive Filtering ... - IJRIT
In general, noise can be reduced using either passive methods or active (Active Noise Control, ANC) techniques. The passive methods works well for high ...

Design and Simulation of Multi-Band Microstrip Patch Antenna ... - IJRIT
The antenna has be designed using High Frequency Substrate Simulator Software (13.0 version). On simulation it could be seen that the antenna was resonant on multiple frequencies where it had good return loss and acceptable VSWR. This antenna can be

Design and Simulation of Adaptive Filtering Algorithms for Active ...
In general, noise can be reduced using either passive methods or active (Active Noise Control, ANC) techniques. The passive methods works well for high ...

CFD Simulation of the Flow Pattern for Drag Reducing Fluids in ...
Joshi (E-mail address: [email protected]). versus Re. The study revealed ..... Durst, F., J. Jovonavic and J. Sender; “LDA Measurement in the. Near Wall Region of ...

1400 TS2 Advanced CFD - Automatisation and Optimisation.pdf ...
There was a problem previewing this document. Retrying... Download. Connect more apps... Try one of the apps below to open or edit this item. 1400 TS2 ...