Sei sulla pagina 1di 19

This article was downloaded by: [University of Winnipeg]

On: 31 August 2014, At: 13:22


Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered
office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK

Chemical Engineering Communications


Publication details, including instructions for authors and
subscription information:
http://www.tandfonline.com/loi/gcec20

RECENT PROGRESS IN HYDRODYNAMICS


OF INVERSE FLUIDIZED BED REACTORS: A
REVIEW
a a a
Naveenji Arun , Ashir Abdul Razack & V. Sivasubramanian
a
Department of Chemical Engineering , NIT Calicut , Kozhikode ,
Kerala , India
Accepted author version posted online: 04 Mar 2013.Published
online: 08 May 2013.

To cite this article: Naveenji Arun , Ashir Abdul Razack & V. Sivasubramanian (2013) RECENT
PROGRESS IN HYDRODYNAMICS OF INVERSE FLUIDIZED BED REACTORS: A REVIEW, Chemical Engineering
Communications, 200:9, 1260-1277, DOI: 10.1080/00986445.2012.744747

To link to this article: http://dx.doi.org/10.1080/00986445.2012.744747

PLEASE SCROLL DOWN FOR ARTICLE

Taylor & Francis makes every effort to ensure the accuracy of all the information (the
“Content”) contained in the publications on our platform. However, Taylor & Francis,
our agents, and our licensors make no representations or warranties whatsoever as to
the accuracy, completeness, or suitability for any purpose of the Content. Any opinions
and views expressed in this publication are the opinions and views of the authors,
and are not the views of or endorsed by Taylor & Francis. The accuracy of the Content
should not be relied upon and should be independently verified with primary sources
of information. Taylor and Francis shall not be liable for any losses, actions, claims,
proceedings, demands, costs, expenses, damages, and other liabilities whatsoever or
howsoever caused arising directly or indirectly in connection with, in relation to or arising
out of the use of the Content.

This article may be used for research, teaching, and private study purposes. Any
substantial or systematic reproduction, redistribution, reselling, loan, sub-licensing,
systematic supply, or distribution in any form to anyone is expressly forbidden. Terms &
Conditions of access and use can be found at http://www.tandfonline.com/page/terms-
and-conditions
Chem. Eng. Comm., 200:1260–1277, 2013
Copyright # Taylor & Francis Group, LLC
ISSN: 0098-6445 print=1563-5201 online
DOI: 10.1080/00986445.2012.744747

Recent Progress in Hydrodynamics of Inverse


Fluidized Bed Reactors: A Review

NAVEENJI ARUN, ASHIR ABDUL RAZACK, AND


V. SIVASUBRAMANIAN
Department of Chemical Engineering, NIT Calicut, Kozhikode,
Kerala, India
Downloaded by [University of Winnipeg] at 13:22 31 August 2014

Hydrodynamic characteristics of the inverse fluidized bed reactor such as phase


holdup, pressure drop, minimum fluidization velocity, liquid circulation velocity,
friction factor, bubble size and bubble rise velocity, and mass-transfer coefficient
along with heat transfer studies, RTD studies, and the influence of these parameters
on the extent of inverse fluidization are reviewed in this work. The importance of
each of these is presented, and certain correlations developed on some hydrodynamic
characteristics are provided. Our work has shown that the majority of investigations
focused on phase holdup, pressure drop, and minimum fluidization velocity, and
studies related to other characteristics such as friction factor, mass-transfer coef-
ficient, and residence time distribution are generally scarce. This review also projects
some of the areas on which future research work can concentrate.

Keywords Hydrodynamics; Inverse fluidized; Correlations; Residence time

Introduction
Inverse bed fluidization incorporates the downward flow of liquid for solid-liquid
fluidization. In solid-liquid fluidization systems, solid particles of density greater
than the liquid are fluidized by the upward movement of the fluid and when the solid
particles have density less than that of the liquid, downward flow is preferred.
Inverse fluidized bed reactors are employed to treat wastewater from various indus-
tries (Bendict et al., 1998; Fan, 1989; Han et al., 2007; Myre and Macchi, 2010; Sokół
et al., 2009). Compared to conventional fluidized bed and packed bed systems, the
advantages of inverse fluidization are efficient control of the process, minimum
solids attrition, less carryover of solid particles (Renganathan & Krishnaiah,
2004a), higher rate of mass transfer, and possibility for re-fluidization (Comte
et al., 1997; Renganathan & Krishnaiah, 2003). Kawalec-Pieternko (2000) reported
that in biological treatment processes, the expansion of the bed accommodates the
cells produced in such a way that prevents problems of clogging that are often
encountered in a packed bed bioreactor. It is reported in the same article that
uniform fluidization without losses of carrier and flexibility in the rate of the oxygen

Address correspondence to Naveenji Arun, Department of Chemical Engineering, Sri


Venkateswara College of Engineering, Chennai 602105, India. E-mail: naveenjiarun@nitc.ac.in

1260
Hydrodynamics of Fluidized Bed Reactors 1261

supply are the advantages of the inverse fluidized bed. Stratification patterns during
fluidization are dependent on liquid velocity and they have an impact on the extent
of inverse fluidization. In the case of inverse fluidization, the stratification patterns
are highly insensitive to liquid velocity in comparison to normal fluidization (Briens
et al., 1999). The most important advantage is the possibility of keeping a bed of
particles separated from a higher shearing that exists in the riser (Nikolov and
Karamanev, 1987).
Successful design and operation of a gas-liquid-solid fluidized bed system
depends on the accurate prediction of fundamental properties of the system, specifi-
cally, the hydrodynamics, the mixing of individual phases, and the heat and mass
transfer properties (Muroyama and Fan, 1985). Fan et al. (1982a) were the first to
study the hydrodynamic characteristics of an inverse fluidized bed using low-density
particles of different diameters and densities. Gas holdup (Han et al., 2003), pressure
Downloaded by [University of Winnipeg] at 13:22 31 August 2014

drop across the bed (Lee et al., 2000), minimum fluidization velocity (Han et al., 2003;
Sivasubramanian, 2009; Sivasubramanian and Velan, 2006; Vijaya Lakshmi et al.,
2000), bubble size and bubble rise velocity (Liu et al., 2007), friction factor (Vijaya
Lakshmi et al., 2000), and mass transfer coefficient (Liu et al., 2007) are some of
the significant hydrodynamic characteristics that have been investigated extensively
in the past. In order to study the hydrodynamic characteristics like pressure drop,
bed voidage, and minimum fluidization velocity, Renganathan and Krishnaiah
(2004b) used the stochastic method of Monte Carlo simulation. The simulation was
validated by measuring the different hydrodynamic variables such as pressure drop,
bed voidage, and minimum fluidization velocity under conditions of steady-state and
unsteady-state bed expansion=contraction. The importance of Monte Carlo simula-
tion, its application, and the prediction of pressure drop, bed voidage, and minimum
fluidization velocity are explained in the same article. Rajasimman et al. (2007)
employed artificial neural network for the modeling of inverse fluidized bed reactors.
Lee et al. (2000) investigated the hydrodynamic transition from fixed bed to fully flui-
dized bed for three-phase inverse fluidization. It has been reported that hysteresis
occurs between fluidization and defluidization, and by the addition of a wetting
agent, occurrence of hysteresis can be substantially reduced. Han et al. (2003) studied
phase holdup and critical fluidization velocity in a three-phase inverse fluidized bed.
It is reported that both gas=liquid-phase holdup and critical fluidization velocity
depend on the density difference between fluid and particles and inverse fluidization
can be made easier by a small density difference. In addition to this, the design and
choice of gas distributor and column size are very important for the determination
of critical fluidization velocity of inverse fluidization.
The main flow regimes during inverse fluidization are dispersed bubble regime,
bubbling regime, and inverse slugging regime. When the gas velocity is high, the
bubbles tends to coalesce, leading to increase in agitation and shear, causing
considerable decrease in liquid-solid phase contact (Han et al., 2000). Bendict et al.
(1998) studied the bed expansion and pressure drop in an inverse fluidized bed
reactor for a two-phase system. It was reported that increase in carboxy methyl cellu-
lose (CMC) concentration and solid density caused a decrease in minimum fluidiza-
tion velocity (Umf). Karamanev and Nikolov (1992) and Fan et al. (1982) developed
semi-empirical models for bed expansion characteristics in two-phase and three-
phase systems respectively. Bendict et al. (1998) evaluated the expansion of bed
height with variation in liquid flow rate from 0 to 1.6 m3=h. The initial bed heights
were maintained at 6  102 m, 17  102 m, and 27  102 m. In all these cases, the
1262 N. Arun et al.

bed height seemed to remain constant until the liquid flow rate reached a value of
0.6 m3=h and then was found to increase steadily up to a liquid flow rate of
1.6 m3=h. However, no information on the upper limit of liquid flow rate for proper
inverse fluidization was reported. The same article grouped the various parameters
that affect bed height in terms of dimensionless numbers given by the following
equation, which predicts data well within 20%:

ðHf =H0 Þ ¼ 2:656 Re1:258


P Ar0:88 ð1Þ

where Hf=H0 indicates the ratio of fluidized bed height to initial bed height, Rep is
modified Reynolds number and Ar represents Archimedes number.
Minimum fluidization velocity in inverse fluidization decreases with increase in
the upward gas velocity (Legile et al., 1992; Zhang et al., 1998). Buffiere and Moletta
Downloaded by [University of Winnipeg] at 13:22 31 August 2014

(1999) studied certain hydrodynamic characteristics of inverse three-phase fluidized


beds using two different types of particles (difference in diameter and density). Two
fluidization mechanisms were observed: a pseudo-fluidized state and a fluidized
state. Increase in gas velocity above a particular value causes the minimal liquid
velocity to reach zero, ensuring fluidization by gas. Moreover, with increase in gas
velocity, the minimal liquid velocity decreases; this state was defined as a ‘‘pseudo-
fluidized state’’ (Legile et al., 1988). For gas velocities lower than the minimum
fluidization velocity, the change in liquid velocity is directly proportional to the gas
velocity. Above minimum fluidization velocity, the liquid velocity increases exponen-
tially with an increase in gas velocity (Kawalec-Pietrenko, 2000).
Choi and Shin (1999) used two different driving forces for fluidization, aeration
and centrifugal force, and studied the hydrodynamic characteristics of the reactors in
both cases. They came to a critical conclusion that a fluidized bed with aeration is
more efficient than a bed where centrifugal force is used as the driving factor. One
major advantage of using aeration as the driving force is that the reactor can be
operated at a low velocity of gas and high particle loading and the gas velocity
for critical fluidization decreases with increase in particle loading. The major factors
that seem to affect fluidization in the case of a centrifugal force–driven reactor are
geometry of reactor spacing and type of impeller used. Kim et al. (2003) studied
liquid axial dispersion and gas-liquid mass transfer in a three-phase inverse fluidized
bed reactor. Axial dispersion was found to increase with gas velocity due to the
increase in turbulence in the fluidized bed. Increase in liquid velocity aids mixing
of fluid elements, resulting in enhanced liquid dispersion, and with increase in liquid
viscosity, the mobility of the liquid element decreases. Therefore, the liquid axial
dispersion decreases with increase in liquid viscosity. According to Kim et al.
(2003), the axial dispersion and mass transfer coefficient are correlated to operating
parameters by Equations (2) and (3) respectively:
 2:26
qs
DZ ¼ 0:157 U0:263
G U0:225
L l0:018
L ð2Þ
ql

 0:854
qs
kL a ¼ 0:229 U0:414
G U0:245
L l0:066
L ð3Þ
ql

The range of validity of Equations (2) and (3) is 0.2 cm=s  UG  0.8 cm=s, 1.0 cm=
s  UL  4.0 cm=s, 0.961 mPa  s  ml  38 mPa  s, and 0.875  qs=ql  0.966.
Hydrodynamics of Fluidized Bed Reactors 1263

Das et al. (2010) determined minimum inverse fluidization velocity for


four different polymeric solids and four different non-Newtonian fluids in two
different columns. In inverse fluidization, very low liquid velocity is required
to fluidize solid articles. Hence, the energy expenditure and sold attrition rate
is low compared to those of ordinary fluidization. Comparison of correlations
developed by other works and Das et al. (2010) is given in Table I. Montastruc
et al. (1991) studied modeling of gas and liquid residence time distribution in
inverse bed bioreactors using a Mellin modification of the Laplace transform-
ation over relevant equations. Alhough the computational fluid dynamics
(CFD) approach is a powerful methodology, the complexity involved is very
high, especially during modeling of bubble breakup or coalescence. More infor-
mation on the modeling algorithm is available in the same article. One major
advantage of this Mellin transformation is the allowance for zero-time solutions
Downloaded by [University of Winnipeg] at 13:22 31 August 2014

for identification analysis. It can be seen from the same article (Figures 5 and 6
of Montastruc et al., 1991) that the experimental and model points seem to be
in good agreement.

Table I. Comparison of correlations from literature

Average
relative
Equation System error (%) Reference
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Rem ¼ ð33:72 þ 0:0408 ArÞ33:7 Solid water; 96.046 Wen and
Classical Yu, 1966
fluidization;
Nonspherical
solids
h i0:38
ðql qs Þ Solid-water; 76.19 Ulaganathan
um ¼ 2:93  103 Ar0:202 ql Solid-LDPE; and
Spherical Krishnaiah,
1996
Rem ¼ 7:68  106 Ar1:513
m
Solid-water; 98.97 Banerjee et al.,
Solid-LDPE; 1999
Spherical,
cylindrical,
and disk
shaped
Rem ¼ 1:139  106 Ar0:041
m
Solid-non- 80.33 Vijaya
Newtonian Lakshmi
CMC et al., 2000
solution;
Solid–LDPE,
PP, Spherical
 0:007
Rem ¼ 1:279 Ar0:549
m
Solid-non- 2.622 Das et al.,
 1:023  0:101 Newtonian 2010
1:857  0:382 dc system
/
dM
1264 N. Arun et al.

The primary aim of this study is to analyze the extent to which research has been
done in the area of inverse bed fluidization and to project the areas that future
researchers can concentrate on. The secondary objective of this study is to give
a brief description of some of the significant parameters that dominate the inverse
fluidization reaction. To the best of our knowledge, no study has investigated or revised
all of the hydrodynamic characteristics pertaining to inverse fluidized bed systems.

Hydrodynamic Characteristics
Phase Holdup
Phase holdup is one of the most significant parameters that determines the effective-
ness of the fluidization process. Many researchers have concentrated on the predic-
tion of the phase holdup for their systems. The phase holdup is found to depend on
Downloaded by [University of Winnipeg] at 13:22 31 August 2014

different parameters like liquid velocity, density of the particle (Cho et al., 2002;
Kim and Kang, 2006; Liu et al., 2007; Shin et al., 2007), gas velocity (Cho et al.,
2002; Lee et al., 2000; Liu et al., 2007), and viscosity of fluid (Shin et al., 2007).
Addition of surfactants lowers the gas velocity required to fluidize the bed and
higher gas holdup can be achieved (Hamdad et al., 2007). It has been reported that
gas holdup increases with gas velocity (Cho et al., 2002; Lee et al., 2000; Liu et al.,
2007; Myre and Macchi, 2010; Shin et al., 2007). Solid holdup is reported to decrease
with increase in gas velocity and liquid velocity (Shin et al., 2007). Liu et al. (2007),
Shin et al. (2007), and Cho et al. (2002) stated that overall gas holdup increased
with superficial liquid velocity. For two-phase systems involving liquid and solid
we have pressure gradient given by:

dP
 ¼ ðes qs þ el ql Þg ð4Þ
dz

es þ el ¼ 1 ð5Þ

Similarly, for a three-phase system, the pressure gradient along the Z-direction
is represented as

dP
 ¼ ðes qs þ el ql þ eg qg Þg ð6Þ
dz
es þ el þ eg ¼ 1 ð7Þ

Whereas the solid phase holdup is given by

Mp
es ¼ p 2
ð8Þ
4 t q s HB
D

From Kim and Kang (2006), the gas phase holdup (eg), gas bubble rising velocity
(UB), frequency of rising bubbles (FB), and liquid axial dispersion coefficient (Dz)
are correlated as
 3:61
qs
eg ¼ 1:18 Ug0:235 Ul0:335 l0:391
l ð9Þ
ql
Hydrodynamics of Fluidized Bed Reactors 1265

 2:13
qs
UB ¼ 0:420 Ug0:114 Ul0:173 l0:132
l ð10Þ
ql

 5:53
qs
FB ¼ 12:4 Ug0:052 Ul0:356 l0:056
l ð11Þ
ql

 2:26
qs
Dz ¼ 0:157 Ug0:263 Ul0:225 l0:018
l ð12Þ
ql

The correlation coefficients are 0.92, 0.95, 0.94, and 0.92 respectively. Shin et al.
(2007) has reported that the variation of liquid holdup with gas velocity is complex
Downloaded by [University of Winnipeg] at 13:22 31 August 2014

as it does not follow a particular pattern of variation. Liquid holdup is found


to increase with increase in liquid velocity and gas velocity (Cho et al., 2002).
At constant gas velocity, the liquid holdup increases, and solid holdup decreases
with increasing liquid velocity due to the expansion of the bed under fluidized
bed conditions (Bandaru et al., 2007). However, in a normal three-phase fluidized
bed, liquid holdup is reported to decrease with increase in gas velocity (Cho et al.,
2002). As the liquid flows downwards, it can interfere with the upward flow of gas
bubbles. Thus the gas holdup is found to increase with increasing liquid velocity.
The same article (Cho et al., 2002) has justified this pattern of variation of phase
holdup with phase velocity. Hydrostatic pressure difference between the riser and
the downcomer is the driving force for liquid circulation in the inverse bed reactors
and this is created mainly by the difference in gas holdup between the riser and
downcomer (Kawalec-Pietrenko, 2000).

Pressure Gradient
Pressure gradient studies are important for the energy and power requirement
calculations in a fluidized bed system. Pressure drop in the reactor can be used to
determine friction factor in the reactor and minimum fluidization velocity in the
cases of two-phase and three-phase fluidization systems. Generally, pressure drop
in the reactor increases with liquid flow rate until a particular value and then is
observed to remain constant over a wide range of liquid flow rates (Bendict et al.,
1998). Pressure gradient can be highly influenced by collision frequency and
collisional particle pressure. Buffiere and Moletta (2000) estimated the frequency
of collisions and the collisional particle pressure and proposed a direct correlation
to estimate the former two parameters as a function of the operating parameters.
The model developed by Buffiere and Moletta is given by:

  b  d
es es
Model ¼ a F UcG ð13Þ
es0 es0

The values of the parameters are: collision frequency (CF), a ¼ 305700, b ¼ 1.01,
c ¼ 1.02, and d ¼ 0.99; and dimensionless particle pressure (PP), a ¼ 423, b ¼ 1,
c ¼ 1.14, and c ¼ 0.75
1266 N. Arun et al.

The empirical correlations developed by Buffiere and Moletta are given by.

 2  
es es
CF ¼ 56; 000 1 UG ðgiven in HzÞ ð14Þ
es0 es0
 1:75  
Pp es es
P0p ¼ 1 2
¼ 422 1  U1:14
G ð15Þ
q u
2 p t
e s0 e s0

Collisional particle pressure can have an influence on the abrasion rate, which
directly affects energy losses (Buffiere and Moletta, 2000). However, this was not
confirmed experimentally. To the best of our knowledge, no other studies analyzed
the interaction between collisional particle pressure and abrasion rate. This could be
a potential field for future researchers to focus on. Lee (2001) reported that as the
Downloaded by [University of Winnipeg] at 13:22 31 August 2014

liquid velocity increases, pressure gradient increases, and it decreases gradually with
increase in liquid velocity beyond a liquid minimum fluidization velocity. This is due
to bed expansion as the fluidization starts. The overall pressure variation in the
vertical direction (z) with correction for frictional pressure gradient is given as
   
dP dP
 ¼ ql g þ  ð16Þ
dz ls dz f ;ls

The frictional pressure gradient in a two-phase (solid-liquid) inverse fluidized bed


is given by
 
dP
 ¼ es ðqs  ql Þg ð17Þ
dz f ;ls
 
For a fixed or moderately expanded bed,  dP dz f ;ls can be expressed by the Ergun
equation (Ergun, 1952) applied to the liquid-solid interaction:

 
dP 150ð1  el Þ2 ll Ul 1:75ð1  el ÞUl2
 ¼  ð18Þ
dz f ;ls /2 dp2 e3l e3l /dp

The final expression for pressure gradient is given by


 
 dP
dz ls 150ð1  el Þ2 ll Ul 1:75ð1  el ÞUl2
¼1  ð19Þ
ql g /2 dp2 e3l ql g e3l /dp g

In order to have a uniform pressure gradient over the particular height interval, voi-
dage, particle properties, and the superficial liquid velocity are assumed to be con-
stant over the height of the bed. In a two-phase system, pressure drop is directly
proportional to initial bed height and CMC concentration (Bendict et al., 1998).
In the case of three-phase systems, the pressure gradient across the height is influ-
enced by density and the void fractions of solid, liquid, and gas phases. Mathemat-
ically, it is represented as
dP
 ¼ ðes qs þ el ql þ eg qg Þg ð20Þ
dz
Hydrodynamics of Fluidized Bed Reactors 1267

The final expression for pressure gradient in a three-phase inverse bed is given by
 
1 dP 150e2 l Ul 1:75es Ul2
 ¼ 1  eg  2 s 3 l  3 ð21Þ
ql g dz / dp2 el ql g el /dp g

The frictional pressure drop Dpf developed in the fully fluidized bed is given by

Dpf ¼ Dpparticles þ Dpgas ð22Þ



ql
Mp g qg 1
Dpparticles ¼ ð23Þ
At
 !
ql
Downloaded by [University of Winnipeg] at 13:22 31 August 2014

qg eg Vtotal g qg 1 q
Dpgas ¼ ¼ qg eg HB g l  1 ð24Þ
At qg

Minimum Fluidization Velocity


Minimum fluidization velocity (Umf) is a basic parameter for the design of a circulat-
ing fluidized bed system (Zhu et al., 2007). The minimum liquid fluidization velocity
is the velocity at which the pressure gradient within the bed is minimum (Cho et al.,
2002; Lee et al., 2000). Renganathan and Krishnaiah (2003) worked on development
of correlations to predict minimum fluidization velocity over a wide range of vari-
ables in two- and three-phase inverse fluidized beds. Cho et al. (2002), determined
liquid minimum fluidization (ULmf) velocity from the pressure gradient in the bed.
For two-phase inverse fluidization, when the pressure drop across the bed is equal
to the net buoyant force per unit area, fluidization starts to occur (Renganathan
and Krishnaiah, 2003). Generally, minimum fluidization velocity (Umf) can be
estimated from the correlation obtained from the data relating the dimensionless
Reynolds number and Archimedes number. Vijaya Lakshmi et al. (2000) reported
that Umf is independent of initial bed height (solid loading) and increased with an
increase in particle diameter and a decrease in density. As reported in the same
article, Umf decreased with increase in concentration of non-Newtonian fluids. In
three-phase systems, the liquid velocity at which the bed fluidizes depends on gas
velocity, and for each gas velocity; the liquid velocity at which the bed will fluidize
differs. The passage of gas reduces the effective density of the liquid phase and the
particles become less buoyant. Minimum liquid fluidization velocity decreases with
increase in gas velocity (Cho et al., 2002; Fan et al., 1982a, 1982b; Ibrahim et al.,
1996), and as the velocity of gas increases, the gas holdup in the bed increases. This
in turn decreases the liquid holdup and, consequently, liquid interstitial velocity
increases. Han et al. (2003) reported that the critical fluidization velocity decreases
with increase in particle loading, a characteristic of three-phase inverse fluidized
beds. Based on the model proposed by Comte et al. (1997), Han et al. (2003)
developed an expression for critical fluidization velocity (UGc) given by
UGc ¼ GRð1  es Þ ð25Þ

where R is the density ratio and G is the bubble swarm velocity.


1268 N. Arun et al.

In normal fluidization from bottom to top, large particle loads are better. The
critical fluidization velocity is more sensitive to the density of particles than to the
surface property of particles. Several models have been proposed for the prediction
of the minimum fluidization velocity in regular three-phase beds. The most suitable
models assume that the effect of the gas bubbles is to displace the liquid, increasing
the velocity of the liquid flowing past the particle surfaces and hence the drag force
of the liquid on the particles (Briens et al., 1999; Zhang et al., 1995). The model from
Zhang et al. (1995), which is easier to adapt to inverse three-phase fluidization,
requires knowledge of the ratio aof the gas holdup egto the bed voidage (1-es), given
by Equation (23).
  
C Ug
a¼ ð26Þ
1  es Ug þ UL
Downloaded by [University of Winnipeg] at 13:22 31 August 2014

where C is a dimensionless empirical constant.

Friction Factor
Friction factor is generally determined from pressure drop data and it is used in the
determination of energy losses (Vijaya Lakshmi et al., 2000). Pressure drop and
friction factor are related by
   
dp DP e3
f ¼ ð27Þ
ql Ul2 L 1e

It is also reported in the same article that the friction factor is highly dependent on
Reynolds number (for NRe < 10) and then it is independent of Reynolds number. The
dimensionless correlation developed by multiple regression analysis relating porous
medium friction factor and porous medium Reynolds number is given by
 
fpm water
¼ 2:31 Re0:626
pm ð28aÞ
 
fpm CMC
¼ 298:4 Re1:24
pm ð28bÞ

where

dp ql Ul
Repm ¼ ð29Þ
lð1  eÞ

From the results of Vijaya Lakshmi et al. (2000), increase in solid loading and
viscosity of the fluid causes an increase in friction factor. Krishnaiah et al. (1993)
proposed an empirical correlation for friction factor in terms of the physical pro-
perties of gas, liquid, solid, and system variables as
 0:47
   
dp 0:126 0:214 1 0:146
f ¼ 18:44 Ar exp 3:39Re g ln  0:014 Ar ð30Þ
L0 e

for Rel > Relmf, with a maximum error of 18%.


Hydrodynamics of Fluidized Bed Reactors 1269

Bubble Size and Bubble Rise Velocity


The properties of the bubbles play a vital role in determining the fluidization
characteristics in inverse fluidized bed systems. Bubble size can be measured by
taking photographs as the fluidization proceeds or by a dual-electrode resistance
probe (Liu et al., 2007). In this article, it is shown that overall bubble diameter
decreased with increase in liquid velocity and solid loading. With increase in gas
velocity, the bubble diameter was reported to increase. Bubble rise velocity increases
with increasing superficial gas velocity and decreased greatly with increase in super-
ficial liquid velocity and solid loading. Similar observations were reported by Kim
and Kang (2006) and Son et al. (2007). Son et al. (2007) reported that the bubble
frequency increases with increasing gas or liquid velocity. The increase of bubble
frequency results in one of the important factors to increase, bubble holdup.
However, the frequency of rising bubbles decreases with increase in liquid viscosity,
Downloaded by [University of Winnipeg] at 13:22 31 August 2014

as the probability of bubble coalescence increases with increasing liquid viscosity


(Son et al., 2007). Bubble size, rising velocity, and frequency of bubbles have been
well correlated based on the concept of the gas drift flux as
 0:446  2:78
UG þ UL qs
LV ¼ 0:117 ðlL Þ0:91 ð31Þ
1  eG qL
   
UG þ UL 0:219 qs 2:89
UB ¼ 0:108 ðlL Þ0:076 ð32Þ
1  eG qL
   
UG þ UL 0:404 qs 6:732
FB ¼ 30:846 ðlL Þ0:002 ð33Þ
1  eG qL

The correlation coefficients of the above equations are 0.91, 0.90, and 0.96
respectively.
Hamdad et al. (2007) reported that the presence of surfactants causes inhibi-
tion of bubble coalescence and surface mobility, resulting in smaller bubbles, lower
bubble rise velocities, and greater gas holdups. However, for larger bubbles, that
is, for larger bubble slip velocities, additives such as sodium chloride (NaCl), sodium
phosphate dibasic (Na2HPO4), iso-amyl alcohol (IAA), and benzoic acid (BA)
reduce gas holdup. A possible explanation was given by Briens et al. (1999), who
stated that the additives stabilize the bubbles, inhibiting the splitting caused by the
turbulent shear forces, which are much stronger in the three-phase fluidized bed.
Additives would have two competing effects since splitting inhibition would reduce
the gas holdup while coalescence inhibition would increase the gas holdup.

Mass-Transfer Coefficient
Mass transfer between phases is a very important parameter for the design and study
of reactors. Liu et al. (2007) developed a new steady-state method for the measure-
ment of dissolved oxygen concentration in liquid. In their studies, the dissolved
oxygen concentration was measured using a covered membrane galvanic oxygen
sensor (WTW Model Cellox 325) connected to an oxygen meter (WTW Model Oxi315i).
Many models, such as the backflow cell model (BFCM), the plug flow model (PFM),
and the axial dispersion model (ADM), can be used to estimate the gas-liquid
1270 N. Arun et al.

mass-transfer coefficient. However, these models are not precise for complex flui-
dized beds with different flow modes in different regions. Hence, Okhotskii (2005)
suggested the use of different flow models for different regions. The BFCM model
was used for the gas distributor region, the downcomer-affected region, and the
gas-slurry separation region, whereas the ADM model was used for the draft tube
region and the annulus region. Generally, kLa value increases with gas velocity Ug
as an increase in gas velocity increases liquid phase turbulence through mixing
(Hamdad et al., 2007; Kim and Kang, 2006). Even though smaller bubbles coalesce
at higher gas velocities, there is a net increase in gas holdup as well as interfacial area
available for transfer (Hamdad et al., 2007). The same article (Hamdad et al., 2007)
reported that an increase in solid loading decreases values of kLa at higher gas
velocities. This is primarily due to decrease in gas holdup and interfacial area.
Similar observations were made by Sánchez et al. (2005). It has been reported that
Downloaded by [University of Winnipeg] at 13:22 31 August 2014

kLa value increases with Ul initially but approaches an asymptotic value with further
increase in Ul (Kim and Kang, 2006). Increase in slip velocity on the bubble caused
by countercurrent liquid flow leads to increase in kL (Linek et al., 2005). For higher
density particles, the value of kLa is more than that for low density particles. By
isotropic turbulence theory, kLa is related as

   3:10  1:74
kL aL Ul qs
St ¼ ¼ 0:719 ð34Þ
Ul Ug þ Ul ql

Nikov and Karamanev (1991) concluded that the density of particles has a strong
influence on the mass transfer rate and the rate of mass transfer is independent of
superficial velocity. Studies by Nikolov et al. (2000) analyzed the influence of gas-
and liquid-phase superficial velocity on mass transfer rate. It was reported in the
same article that superficial gas velocity had a significant positive impact on mass
transfer rate in comparison to the superficial liquid velocity. The statement by Nikov
and Karamanev (1991) seems to contradict the conclusions by Nikolov et al. (2000).

Heat Transfer Characteristics


Heat transfer studies in inverse fluidization are comparatively scarce. Regulating
temperature during inverse fluidization is crucial; it can affect the mass transfer rate
since viscosity is related to the temperature of the fluid system. Myri and Macchi
(2010) studied three-phase inverse fluidization and evaluated the average surface-to-
bed heat transfer coefficient over a wide range of superficial gas velocities and bed
expansions. Heat transfer characteristics for two- and three-phase inverse fluidized
beds were studied by Cho et al. (2002). The heat transfer rate can be affected by
gas and liquid superficial velocities, solids loading, and the presence of surfactants.
Increase in bubble coalescence promoted by increase in superficial gas and liquid
velocity caused an increase in heat transfer coefficient. Based on the conclusions
of Myre and Macchi (2010), the presence of surfactant affected gas holdup and
caused a decrease in heat transfer coefficient by decreasing the amount and size of
bubbles. Heat transfer coefficient can be correlated to gas holdup based on the
equations of Behkish et al. (2006) and Son et al. (2007). Temperature can influence
the mass transfer in the two- and three-phase inverse fluidized beds because the
liquid properties such as viscosity or fluidity are strong functions of temperature.
Hydrodynamics of Fluidized Bed Reactors 1271

Therefore, for the commercialization of inverse bed reactors, heat transfer studies
pertaining to the system are highly essential. Cho et al. (2002) focused on determining
the effects of gas and liquid velocities and particle density on the immersed heater-to-
bed heat transfer coefficient. Moreover, the heat transfer coefficient for the immersed
heater-to-bed system was well correlated in terms of dimensionless numbers as given
below. With the Nusselt number, the heat transfer coefficients are related as

  
h1 dp ð1  es Þ CP lL dp ql Ul 1:282
Nu1 ¼ ¼ 0:001 ð35aÞ
kL es kL ll es

  
h1 dp ð1  es Þ CP lL dp ql Ug 0:944
Nu2 ¼ ¼ 0:084 ð35bÞ
kL es kL ll es
Downloaded by [University of Winnipeg] at 13:22 31 August 2014

  
h3 dp ð1  es Þ CP lL dp ql ðUg þ Ul Þ 0:810
Nu3 ¼ ¼ 0:050 ð35cÞ
kL es kL ll es

The correlation coefficients of the equations are 0.95, 0.91, and 0.94, respect-
ively. Nu1, Nu2, and Nu3 are the Nusselt numbers for two-phase (liquid-solid)
systems, three-phase with batch condition (UL ¼ 0), and three-phase with UL 6¼ 0
respectively. Cho et al. (2002) have given detailed explanations of the variation of
heat transfer coefficient with phase velocity and the reasons for such variations.
The heat transfer coefficients for two- and three-phase inverse beds for high-density
particles are higher than that of beds with low-density particles. Cho et al. (2002) and
Myre and Macchi (2010) have stated that coefficient value increases with gas velocity
due to the vigorous bubbling at high gas velocity. With high liquid velocity, the solid
holdup decreases and the heat transfer coefficient exhibits a maximum value. In
a conventional bed, the coefficient increases with liquid velocity as the turbulence
and particle mobility increases. Son et al. (2007) studied the influence of parameters
like gas and liquid velocity, liquid viscosity, particle density, and bed porosity on
the heat transfer coefficient.

Residence Time Distribution (RTD) Studies


Renganathan and Krishnaiah (2003) studied the residence time distribution the
using pulse tracer technique and the deconvolution method of analysis in two-phase
liquid-solid inverse fluidized bed reactors. They determined the Peclet number and
the dispersion coefficient. The axial dispersion model impulse response function
f ðt; PeÞ (Levenspiel, 2001) is given by

 1 " #
1 Pe 2 Pe ðh  1Þ2
f ðt; PeÞ ¼ exp  ð36Þ
2 ph3 4 h

where h ¼ tt is the dimensionless residence time of the system.


With open boundary conditions, the RTD of the system obtained from the axial
dispersion model (Levenspiel, 2001) is expressed as
1272 N. Arun et al.

  " #
1 Pe 1=2 Pe ðh  1Þ2
EðhÞ ¼ exp  ð37Þ
2 ph 4 h

By applying the deconvolution procedure, the liquid phase axial dispersion


coefficient was determined using the equation

UlDz
Dl ¼ ð38Þ
Pee

where Dz is the distance between the two measuring points. The values for D1
obtained by Renganathan and Krishnaiah (2003) are in the range of 0.1–100 cm2=s.
This is similar to the results obtained by previous researchers on normal fluidized
Downloaded by [University of Winnipeg] at 13:22 31 August 2014

beds. Moreover, the effect of liquid velocity, static bed height, and Archimedes
number were studied in the same work. It was stated that the axial dispersion
coefficient remains constant and Peclet number increases with increase in static
bed height. With increase in Archimedes number, the axial dispersion coefficient is
reported to increase owing to the fact that minimum fluidization velocity and actual
liquid velocity also increase. The axial dispersion coefficient is reported to increase
with increase in liquid velocity as the turbulence in the bed is increased. The
correlation developed by Renganathan and Krishnaiah (2003) for determining axial
dispersion coefficient is given by (RMS of 28%):

 1:73
4 0:66 Re
D1 ¼ 1:48  10 Ar ð39Þ
Remf

This equation is valid for

17:6 < Ar < 1:47107 ð40aÞ

0:036 < Re < 1267 ð40bÞ

From the Wen and Yu equation (Wen and Yu, 1996):

qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Remf ¼ ð33:7Þ2 þ 0:0408 Ar  33:7 ð41Þ

This Wen and Yu equation has been validated by the same authors for determining
the minimum fluidization velocity in a two-phase liquid-solid inverse fluidized bed.
The value of particle dispersion coefficient DP decreases gradually with increasing
liquid viscosity. With increase in liquid viscosity in the inverse bed, the mobility of
fluidized particles and fluctuating frequency could decrease, minimizing the disper-
sion behavior of particles (Lee et al., 2007). It has been pointed out that the bubble
size in a three-phase inverse fluidized bed increases with increase in viscosity of
the liquid medium owing to a decrease of bubble-breaking potential of particles in
the beds (Son et al., 2007).
Hydrodynamics of Fluidized Bed Reactors 1273

Conclusions
Inverse fluidization systems have been of considerably less importance in the past
but seem to be gaining importance owing to their varied industrial applications.
Studies pertaining to inverse fluidized beds and their characteristics have been
reviewed. Presently, most of the hydrodynamic characteristics have been studied
by many researchers worldwide. Pressure drop, minimum fluidization velocity, phase
holdup, and mass-transfer coefficient are some of the most significant characteristics,
and many studies have focused on them. However, some have focused on RTD, bed
dispersion, and many more topics, adding value to the scope of inverse fluidized bed
reactors. The results for different characteristics obtained by various researchers are
similar in most cases. It can be concluded from this review that RTD, heat transfer,
friction factor, and bed expansion are some of the subjects that need more attention
in future research. Maximum exploration in these areas is essential for the novel
Downloaded by [University of Winnipeg] at 13:22 31 August 2014

design and application of inverse fluidized beds at the commercial scale.

Nomenclature
At cross-sectional area, m2
a interfacial area per unit volume of liquid, m2  m3
Arm modified Archimedes number, dimensionless
Arpl power law Archimedes number, dimensionless
C empirical constant, dimensionless
Cp heat capacity of the liquid phase, J  kg1K1
Dl, Dz liquid axial dispersion coefficient, cm2  s1
Dt diameter of the column, m
dp particle diameter, m
E(h) dimensionless exit age distribution
FB frequency of rising bubble, s1
f friction factor, dimensionless
fpm porous medium friction factor, dimensionless
G bubble swarm velocity, cm  s1
g acceleration due to gravity, m  s2
HB bed height, m
h1 heat transfer coefficient for two-phase system, W  m2 K1
h2 heat transfer coefficient for three-phase batch system, W  m2 K1
h3 heat transfer coefficient for three-phase system, W  m2 K1
hD height of the dispersion, m
K consistency index in power-law model, Pa  sn
KB frictional loss coefficient, dimensionless
kL liquid-side mass-transfer coefficient, m3  m2s1
kLa volumetric gas-liquid mass-transfer coefficient, s1
L length of column, m
LV bubble size, m
Mp particle inventory, kg
Nu1 Nusselt number for two-phase system, dimensionless
Nu2 Nusselt number for three-phase batch system, dimensionless
Nu3 Nusselt number for three-phase system, dimensionless
n flow index in power-law model, dimensionless
1274 N. Arun et al.

DP pressure drop across bed, Pa


Pe Peclet number, dimensionless
Dpf frictional pressure drop, Pa
Dpgas pressure drop of gas, Pa
Dpparticles
 pressure drop due to particles, Pa
 dP
dz
overall pressure drop in liquid-solid system, Pa  m1
 dPls
 dz f ;ls frictional pressure gradient due to liquid-solid interaction, Pa  m1
Re, NRe Reynolds number, dimensionless
Remf Reynolds number at minimum fluidization condition, dimensionless
(Repl)CMC power-law Reynolds number, dimensionless
(Rep)water particle Reynolds number, dimensionless
Repm porous medium Reynolds number, dimensionless
St Stanton number, dimensionless
Downloaded by [University of Winnipeg] at 13:22 31 August 2014

t time, s
tc circulation time, s
t mean residence time, s
UB bubble rise velocity, m  s1
UGc critical gas velocity, m  s1
Ug superficial gas velocity, m  s1
Ul superficial liquid velocity, m  s1
ULmf liquid minimum fluidization velocity, m  s1
Umf minimum fluidization velocity, m  s1
VL volume of liquid in reactor, m3
Vtotal total volume in the bed, m3
Dz axial distance between pressure measurement, m
Greek letters
a solids free gas holdup, dimensionless
e porosity of bed, dimensionless
eg gas holdup, dimensionless
el liquid holdup, dimensionless
es solid holdup, dimensionless
ml liquid viscosity, kg  m1s1
qg density of gas, kg  m3
ql density of liquid, kg  m3
qs density of solid, kg  m3
/ particle sphericity, dimensionless

References
Bandaru, K. S. V. S. R., Murthy, D. V. S., and Krishnaiah, K. (2007). Some hydrodynamic
aspects of 3-phase inverse fluidized bed, China Particuology, 5, 351–356.
Banerjee, J., Basu, J. K., and Ganguly, U. P. (1999). Some studies on the hydrodynamics of
reverse fluidization velocities, Ind. Chem. Eng., 41(1), 35–38.
Behkish, A., Lemoine, R., Oukaci, R., and Morsi, B. I. (2006). Novel correlations for gas
holdup in large-scale slurry bubble column reactors operating under elevated pressures
and temperatures, Chem. Eng. J., 115, 157–171.
Bendict, R. J. F., Kumaresan, G., and Velan, M. (1998). Bed expansion and pressure drop
studies in a liquid-solid inverse fluidized bed reactor, Bioprocess Eng., 19, 137–142.
Hydrodynamics of Fluidized Bed Reactors 1275

Briens, C. L., Ibrahim, Y. A. A., Margarities, A., and Bergougnou, M. A. (1999). Effect of
coalescence inhibitors on the performance of three-phase inverse fluidized bed columns,
Chem. Eng. Sci., 54, 4975–4980.
Buffiere, P, and Moletta, R. (1999). Some hydrodynamic characteristics of inverse three phase
fluidized-bed reactors, Chem. Eng. Sci., 54, 1233–1242.
Buffiere, P, and Moletta, R. (2000). Collision frequency and collisional particle pressure in
three-phase fluidized beds, Chem. Eng. Sci., 55, 5555–5563.
Cho, Y. J., Park, H. Y., Kim, S. W., and Kang, Y. (2002). Heat transfer and hydrodynamics in
two- and three-phase inverse fluidized beds, Ind. Eng. Chem. Res., 41, 2058–2063.
Choi, H. S., and Shin, M. (1999). Hydrodynamics study of two different inverse fluidized
reactors for the application of wastewater treatment, Korean J. Chem. Eng., 16(5), 670–676.
Comte, M. P., Bastoul, D., Hebrarel, G., Roustan, M., and Lazarova, V. (1997). Hydrodyn-
amics of a three-phase fluidized bed—The inverse turbulent bed, Chem. Eng. Sci., 52,
3971–3977.
Downloaded by [University of Winnipeg] at 13:22 31 August 2014

Das, B., Ganguly, U. P., and Das, S. K. (2010). Inverse fluidization using non-Newtonian
liquids, Chem. Eng. Process., 49, 1169–1175.
Ergun, S. (1952). Fluid flow through packed columns, Chem. Eng. Prog., 48, 89–94.
Fan, L. S. (1989). Gas–Liquid–Solid Fluidization Engineering, Butterworths, Boston.
Fan, L. S., Muroyama, K., and Chern, S. H. (1982a). Hydrodynamic characteristics of inverse
fluidization in liquid–solid and gas–liquid–solid systems, Chem. Eng. J., 24, 143–150.
Fan, L. S., Muroyama, K., and Chern, S. H. (1982b). Some remarks on hydrodynamics of
inverse gas–liquid–solid fluidization, Chem. Eng. Sci., 37, 1570–1572.
Hamdad, I., Hashmi, S., Rossi, D., and Macchi, A. (2007). Oxygen transfer and hydro-
dynamics in three-phase inverse fluidized beds, Chem. Eng. Sci., 62, 7399–7405.
Han, S. J., Tan, R. B. H., and Loh, K. C. (2000). Hydrodynamic behavior in a new gas-liquid-
solid inverse fluidization airlift bioreactor, Trans Inst. Chem. Eng., 78C, 207–215.
Han, H. D., Lee, W., Kang, K. Y., Lee, K. J., Suk, C. H., Kang, Y., and Done, K. S. (2003).
Phase hold-up and critical fluidization velocity in a three-phase inverse fluidized bed,
Korean J. Chem. Eng., 20(1), 163–168.
Ibrahim, Y. A. A., Briens, C. L, Margaritis, A., and Bergongnou, M. A. (1996). Hydrodynamic
characteristics of a three-phase inverse fluidized bed column, AIChE J., 42, 1889–1900.
Karamanev, D. G., and Nikolov, L. N. (1992). Bed expansion of liquid–solid inverse
fluidization, AIChE J., 38, 1916–1922.
Kawalec-Pietrenko, B. (2000). Liquid circulation velocity in the inverse fluidized bed airlift
reactor, Bioprocess Eng., 23, 397–402.
Kim, S. D., and Kang, Y. (2006). Hydrodynamics, heat and mass transfer in inverse and
circulating three-phase fluidized-bed reactors for waste water treatment, Stud. Surf.
Sci. Catal., 159, 103–108.
Kim, S. W., Kim, H. T., Song, P. S., Kang, Y., and Kim, S. D. (2003). Liquid dispersion
and gas-liquid mass transfer in three-phase inverse fluidized beds, Can. J. Chem. Eng.,
81, 621–625.
Krishnaiah, K., Guru, S., and Sekar, V. (1993). Hydrodynamic studies on inverse gas-liquid-
solid fluidization, Chem. Eng. J., 51, 109–112.
Lee, D. H. (2001). Transition velocity and bed expansion of two-phase (liquid-solid)
fluidization systems, Korean J. Chem. Eng., 18(3), 347–351.
Lee, D. H., Epstein, N., and John, R. G. (2000). Hydrodynamic transition from fixed to fully
fluidized beds for three-phase inverse fluidization, Korean J. Chem. Eng., 17(6), 684–690.
Lee, K. I., Son, S. M., Kim, U. Y., Kang, Y., Kang, S. H., Kim, S. D., Lee, J. K., Seo, Y. C.,
and Kim, W. H. (2007). Particle dispersion in viscous three-phase inverse fluidized beds,
Chem. Eng. Sci., 62, 7060–7067.
Legile, P., Menard, G., Laurent, C., Thomas, D., and Bernis, A. (1988). Contribution à l’etude
hydrodynamique d’un lit fluidisé triphasique fonctionnant à contren courant, Entropie,
143–144, 23–31.
1276 N. Arun et al.

Legile, P., Menard, G., Laurent, C., Thomas, D., and Bernis, A. (1992). Contribution to the
study of an inverse three-phase fluidized bed operating counter-currently, Int. Chem. Eng.
J., 32, 41.
Levenspiel, O. (2001). Chemical Reaction Engineering, John Wiley & Sons, New York.
Linek, V., Kordac, M., and Moucha, T. (2005). Mechanism of mass transfer from bubbles in
dispersions. Part II: Mass transfer coefficients in stirred gas–liquid reactor and bubble
column, Chem. Eng. Process., 44, 121–130.
Liu, M., Lu, C., Shi, M., Ge, B., and Huang, J. (2007). Hydrodynamics and mass transfer in
a modified three-phase airlift loop reactor, Pet. Sci., 4, 91–96.
Montastruc, L., Brienne, J. P., and Nikov, I. (2009). Modeling of residence time distribution:
Application to a three-phase inverse fluidized bed based on a Mellin transform, Chem.
Eng. J., 148, 139–144.
Muroyama, K., and Fan, L. S. (1985). Fundamentals of gas-liquid-solid fluidization, AIChE
J., 31(1), 1–34.
Downloaded by [University of Winnipeg] at 13:22 31 August 2014

Myre, D., and Macchi, A. (2010). Heat transfer and bubble dynamics in a three-phase inverse
fluidized bed, Chem. Eng. Process., 49, 523–529.
Nikolov, L. N., and Karamanev, D. G. (1987). Experimental study of the inverse fluidized bed
biofilm reactor, Can. J. Chem. Eng., 65, 214–219.
Nikolov, V., Fagar, I., and Nikov, I. (2000). Gas-liquid mass transfer in bioreactor with
three-phase inverse fluidized bed, Bioprocess Eng., 23, 427–429.
Nikov, I., and Karamanev, D. (1991). Liquid-solid mass transfer in inverse fluidized bed,
AIChE J., 37, 781–784.
Okhotskii, V. B. (2005). Hydrodynamic processes in fluidization, Theor. Found. Chem. Eng.,
39(5), 542–547.
Rajasimman, M., Govindarajan, L., and Karthikeyan, C. (2007). Artificial neural network
modeling of an inverse fluidized bed bioreactor, J. Appl. Sci. Environ. Manage., 11(2),
65–69.
Renganathan, T., and Krishnaiah, K. (2003). Prediction of minimum fluidization velocity
in two and three phase inverse fluidized beds, Can. J. Chem. Eng., 81, 853–860.
Renganathan, T., and Krishnaiah, K. (2004a). Liquid phase mixing in 2-phase liquid–solid
inverse fluidized bed, Chem. Eng. J., 98, 213–218.
Renganathan, T., and Krishnaiah, K. (2004b). Stochastic simulation of hydrodynamics of
a liquid-solid inverse fluidized bed, Ind. Eng. Chem. Res., 43, 4405–4412.
Sánchez, O., Michaud, S., Escudié, R., Delgenès, J. P., and Bernet, N. (2005). Liquid mixing
and gas–liquid mass transfer in a three-phase inverse turbulent bed reactor, Chem. Eng.
J., 114, 1–7.
Shin, I. S., Son, S. M., Kang, Y., Kang, S. H., and Kim, S. D. (2007). Phase holdup charac-
teristics of viscous three-phase inverse fluidized beds, J. Ind. Eng. Chem., 13(6), 971–978.
Sivasubramanian, V. (2009). Gas-liquid mass transfer in three-phase inverse fluidized bed
reactor with Newtonian and non-Newtonian fluids, Asia-Pac. J. Chem. Eng., 5, 361–368.
Sivasubramanian, V., and Velan, M. (2006). Modified gas-perturbed liquid model (GLPM)
for minimum fluidization velocity in a two-phase inverse fluidized bed reactor with
Newtonian and non-Newtonian systems, J. Chem. Technol. Biotechnol., 81, 1232–1238.
Sokół, W., Ambaw, A., and Woldeyes, B. (2009). Biological wastewater treatment in the
inverse fluidized bed reactor, Chem. Eng. J., 150, 63–68.
Son, M. S., Kang, S. H., Kim, U. K., Kang, Y., and Kim, S. D. (2007). Bubble properties
in three-phase inverse fluidized beds with viscous liquid medium, Chem. Eng. Process.,
46, 736–741.
Ulaganathan, K., and Krishnaiah, K. (1996). Hydrodynamic characteristics of two-phase
inverse fluidized bed, Bioprocess Eng., 15, 159–164.
Vijaya Lakshmi, A. C., Balamurugan, M., Sivakumar, M., Newton Samuel, T., and Velan, M.
(2000). Minimum fluidization velocity and friction factor in a liquid-solid inverse
fluidized bed reactor, BioProcess Eng., 22, 461–466.
Hydrodynamics of Fluidized Bed Reactors 1277

Wen, C. Y., and Yu, Y. H. (1966). A generalized method for predicting the minimum
fluidization velocity, AIChE J., 12, 610–612.
Zhang, J. P., Epstein, N., Grace, J. R., and Zhu, J. (1995). The pseudo-fluid model applied to
three-phase fluidization, Trans. Inst. Chem. Eng., 73A, 347–353.
Zhang, J.-P., Epstein, N., and Grace, J. R. (1998). Minimum fluidization velocities for
gas-liquid-solid three-phase systems, Powder Technol., 100, 113–118.
Zhu, Z., Na, Y., and Lu, Q. (2007). Effect of pressure on minimum fluidization velocity,
J. Therm. Sci., 16(3), 264–269.
Downloaded by [University of Winnipeg] at 13:22 31 August 2014

Potrebbero piacerti anche