You are on page 1of 26

Solvent Extraction and Ion Exchange

ISSN: (Print) (Online) Journal homepage: https://www.tandfonline.com/loi/lsei20

Computational Fluid Dynamics Modelling to


Predict Axial Dispersion in Pulsatile Liquid-liquid
Two-phase Flow in Pulsed Sieve Plate Columns

Nirvik Sen , K.K. Singh , A. W. Patwardhan & K.T. Shenoy

To cite this article: Nirvik Sen , K.K. Singh , A. W. Patwardhan & K.T. Shenoy (2020):
Computational Fluid Dynamics Modelling to Predict Axial Dispersion in Pulsatile Liquid-liquid
Two-phase Flow in Pulsed Sieve Plate Columns, Solvent Extraction and Ion Exchange, DOI:
10.1080/07366299.2020.1840030

To link to this article: https://doi.org/10.1080/07366299.2020.1840030

Published online: 24 Nov 2020.

Submit your article to this journal

Article views: 8

View related articles

View Crossmark data

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=lsei20
SOLVENT EXTRACTION AND ION EXCHANGE
https://doi.org/10.1080/07366299.2020.1840030

Computational Fluid Dynamics Modelling to Predict Axial


Dispersion in Pulsatile Liquid-liquid Two-phase Flow in
Pulsed Sieve Plate Columns
Nirvik Sena,b, K.K. Singh a,b
, A. W. Patwardhan c
, and K.T. Shenoyb
a
Homi Bhabha National Institute, Mumbai, India; bChemical Engineering Division, Bhabha Atomic
Research Centre, Trombay, Mumbai, India; cInstitute of Chemical Technology, Matunga, Mumbai, India

ABSTRACT KEYWORDS
Computational Fluid Dynamics (CFD) is used to predict contin­ Axial dispersion coefficient;
uous phase axial dispersion coefficient for liquid-liquid two- CFD; population balance;
phase flow in a pulsed-sieve plate column (PSPC). The CFD pulsed sieve plate column;
model is based on Euler-Euler model in which the drop diameter sauter mean drop diameter
used in the interphase momentum exchange term is obtained
by coupled solution of the population balance equations.
Species transport is additionally solved for the continuous
phase to obtain its residence time distribution (RTD) curve
which is further processed to estimate axial dispersion coeffi­
cient of the continuous phase. Experimental studies are also
conducted in a 3-inch diameter PSPC to obtain axial dispersion
coefficient in the continuous phase for different operating con­
ditions. The continuous phase is water. 30% (v/v) tributyl phos­
phate mixed in dodecane is the dispersed phase. Validation of
the CFD model is done with the experimental data of axial
dispersion coefficient of the continuous phase. The average
absolute relative error between estimated and predicted axial
dispersion coefficient of continuous phase is less than 3%. The
validated model is used to understand two-phase hydrody­
namics in the column.

Introduction
Liquid–liquid extraction is used widely in the chemical process industry. It is
a commonly used separation process in hydrometallurgical industries.[1–4]
Mixer-settler, rotary agitated column, annular centrifugal extractor, Kuehni
column, and air-pulsed column are some of the equipment which are used
for carrying out liquid-liquid extraction.[5–8] Among these equipment, air-
pulsed columns, by virtue of being maintenance-free, find several unique
applications. In an air-pulsed sieve-plate column (PSPC), the pneumatically
driven pulsing action drives the counter-current flow of two immiscible
liquid phases. Air pulsing can help achieve high throughput and high separa­
tion efficiency.[9–11]

CONTACT K.K. Singh kksingh@barc.gov.in Homi Bhabha National Institute, Mumbai, India;
A. W. Patwardhan aw.patwardhan@ictmumbai.edu.in Institute of Chemical Technology, Matunga, Mumbai,
India.
© 2020 Taylor & Francis Group, LLC
2 N. SEN ET AL.

Extensive experimental research has been reported on hydrodynamics of


PSPCs. Yadav and Patwardhan[12] gave a comprehensive review of the experi­
mental work reported on PSPCs. They also compared reported correlations for
estimating hydrodynamic variables such as dispersed phase holdup, drop size,
and flooding conditions. The authors also summarized published literature on
mass transfer in PSPCs. Axial dispersion model has been widely employed to
model mass transfer in columnar contractors including PSPCs.[13] In this
model, axial dispersion coefficient characterizes the degree of back-mixing in
the column and affects the driving force for mass transfer. A value of axial
dispersion that is too high reduces the driving force thus having a detrimental
effect on mass transfer. It has been reported that predicting mass transfer
performance by neglecting axial dispersion can lead to significant over-
prediction of mass transfer.[14,15] Axial dispersion in the dispersed phase
does not affect mass transfer in PSPCs significantly. Thus, primarily the axial
dispersion in the continuous phase is considered important .[16,17]
Continuous phase axial dispersion or back-mixing in PSPCs has been
studied by various researchers. Prvcic and co-workers reported back-mixing
in continuous phase for three different columns having diameter equal to 72,
152 and 300 mm.[18] Both single-phase (water) and two-phase (kerosene-
water) experiments were reported. Axial dispersion in single-phase flow was
reported to increase with increase in column diameter while axial dispersion in
two-phase flow was found to be independent of column diameter. Baird
experimentally measured axial dispersion coefficients in a PSPC having
150 mm diameter.[19] They used kerosene-water system and reported increase
in axial dispersion coefficient with an increase in dispersed phase velocity and
pulsing velocity. Yu and Kim used a 102 mm diameter column to study axial
dispersion in continuous phase.[20] The axial dispersion coefficient was
reported to increase with both pulsing amplitude and frequency, but it
reduced with reduction in plate spacing. Rao and co-workers studied contin­
uous phase axial dispersion in a 2-inch diameter PSPC.[16] The phase system
used was water and 30% TBP-kerosene. Axial dispersion coefficient of the
continuous phase was reported to increase with increase in pulsing velocity
while for constant pulsing velocity it increased with increase in continuous
phase superficial velocity. Axial dispersion coefficient initially increased with
an increase in dispersed phase velocity but at higher values of dispersed phase
velocity it did not increase further. Axial dispersion coefficient was reported to
reduce with increase in plate spacing. Niebuhr and Vogelpohl studied axial
dispersion in both continuous and dispersed phases in an 80 mm diameter
PSPC.[21] They reported that axial dispersion coefficient was strongly influ­
enced by behaviour/flow pattern of the dispersed phase. Noh and Kim
reported continuous phase axial dispersion in a 108 mm diameter PSPC.[22]
An acid-base reaction was used as the indicator. Axial dispersion was observed
to increase proportionally with the product of the square of pulsing amplitude
SOLVENT EXTRACTION AND ION EXCHANGE 3

and pulsing frequency. The authors reported that axial dispersion decreased
as percent opening area of the plates was increased. It also reduced with
increase in plate spacing. The plate thickness and flow rates of the liquid
phases were found to have an insignificant effect on axial dispersion.
Several correlations for prediction of axial dispersion coefficient have also
been reported. A generalized correlation for estimating axial dispersion coeffi­
cient in continuous phase was reported by Tung and Luecke.[23] The correla­
tion was obtained by regressing published data for PSPCs, and accounted for
the effect of column geometry. A similar correlation was proposed by
Srinikethan and co-workers taking into account the effects of pulsing ampli­
tude, pulsing frequency, continuous phase velocity and plate geometry (hole
diameter, interplate spacing) on axial dispersion coefficient.[24] The authors
provided correlations for both mixer-settler and emulsion regime of column
operation. Kumar and Hartland compiled a large data set from reported
studies and proposed a unified correlation for estimating continuous phase
axial dispersion coefficient in PSPCs.[25] The correlation could be used for
different liquid-liquid systems and over a wide range of operating conditions.
CFD modelling of axial dispersion in single-phase flow in PSPC has been
reported. Kolhe and co-workers reported a 3D CFD model aimed at estimat­
ing axial dispersion for single-phase flow in a PSPC having 3-inch diameter.[26]
They validated their model with the experimental data of axial dispersion
coefficient. They reported an increase in single-phase axial dispersion with
increase in superficial flow velocity as well as pulsing velocity. Xiaojin and
Guangsheng reported a 3D model to predict axial dispersion coefficient in
coalescence-dispersion pulsed-sieve-plate columns (CDPSPCs).[27] The
authors compared axial dispersion in two different designs of CDPSPC. In
our previous work, we reported a single-phase CFD model to predict axial
dispersion in a PSPC having 3-inch diameter.[28] The model was validated with
reported experimental results. The effect of plate geometry on axial dispersion
coefficient was also reported.
Several studies on two-phase CFD modelling of air pulsed columns have
been reported, but most of them are on pulsed columns having disc and
doughnut shaped internals.[14,29–34] Studies on CFD modelling of two-phase
flow in PSPCs are very few. Yadav and Patwardhan simulated two-phase flow in
pulsed and non-pulsed sieve tray column of 0.05 m diameter.[35] The effects of
pulsing on operating regimes, holdup of the dispersed phase, and column
hydrodynamics were reported. However, the sieve plates used in their study
had a downcomer. The hydrodyanmics in a column having plates with down­
comers may be quite different from the columns having sieve plates without
a downcomer. Din and coworkers reported a two-dimensional model to simu­
late two-phase flow in a PSPC.[36] The authors used Euler-Euler two-fluid
model and mixture k-ε model of turbulence. The section consisting of sieve
plates was modelled as a porous medium. Predicted value of dispersed phase
4 N. SEN ET AL.

holdup was about 70% of the experimentally determined value. Recently we


reported CFD modelling of two-phase flow in a 3-inch diameter PSPC.[37] A 2D
model was reported and validated with experimental data. The Euler-Euler two-
fluid model and mixture k-ε model of turbulence were used to model the two-
phase turbulent flow field. A suitable drag model for simulating pulsatile two-
phase flow in PSPC was identified. However, a prior knowledge of representa­
tive drop diameter was a requirement in this model, which was its limitation.
Thereafter, we reported an approach in which the representative drop diameter
to be used in the CFD model could be estimated from a suitable correlation.[38]
To rid of the assumption of monodispersed droplets, we recently reported for
the first time a CFD-PB model of PSPCs to simultaneously predict dispersed
phase holdup and Sauter mean drop diameter.[39] The model was validated with
reported experimental data of a 3-inch diameter PSPC. The CFD-PB model
does away with the assumption of invariant drop size by considering spatially
and temporally varying drop size distribution inside the column. The method
of classes was used to solve the population balance equations while Euler-Euler
two-fluid approach was used to model the two-phase flow.
Numerical modelling to predict axial dispersion coefficient in PSPCs in
single-phase pulsatile flow has been reported in several previous studies.[26–28]
In a very recent work, two-phase CFD modelling to predict dispersed phase
holdup and axial dispersion in standard-pulsed sieve plate column and hybrid-
pulsed sieve plate column was reported.[40] This study, to the best of our
knowledge, is the only study to date on numerical prediction of axial disper­
sion coefficient in two-phase flow in PSPC.
In the present work, we report for the first time, a two-phase CFD model
coupled with a population balance model to predict continuous phase axial
dispersion in pulsed sieve plate columns. Euler-Euler model is used to simulate
liquid–liquid two-phase dispersed flow. A method of classes is used for
population balance modelling. Mixture k-ε model is used to model the turbu­
lence inside the column. CFD-based virtual tracer experiments are carried out
to obtain the residence time distribution of the continuous phase. Experiments
are performed in a 3-inch diameter PSPC to obtain axial dispersion coefficient
in the continuous phase. Model predictions are validated with experimental
results on axial dispersion coefficient in the continuous phase. Thus, the model
reported in this work predicts all hydrodynamic parameters (i.e. dispersed
phase holdup, Sauter mean drop diameter, axial dispersion coefficient)
required for estimation of mass transfer performance of a PSPC by using an
axial dispersion model.

Experimental setup
Figure 1 shows the schematic diagram of the experimental setup. The setup
has a cylindrical glass column (active section) having 0.076 m diameter and
SOLVENT EXTRACTION AND ION EXCHANGE 5

Figure 1. Schematic diagram of the experimental setup.

0.5 m height. There are two disengagement sections, one each at the top end
and the bottom end of the glass column for density difference driven phase
separation. The aqueous and the organic phases are fed to the column from
the feed tanks by using two centrifugal pumps. Volumetric flow rates of the
phases can be varied by using control valves and the rotameters. A 3-way
valve and compressed air are used to provide air pulsing in the pulse leg.
Water is used as the continuous phase in the experiments. Tributyl phos­
phate (TBP) dissolved in dodecane is the dispersed phase. The proportion
of TBP in the dispersed phase is 30% by volume. A typical experiment
consists of initially filling the column with continuous phase and starting
the pulsing at the desired pulsing velocity (pulsing frequency maintained at
1 Hz). Then, the flow of the continuous phase is started. Thereafter, flow of
the dispersed phase is started. The desired flow rates of the continuous and
dispersed phases are fixed with the help of manual control valves and
rotameters. KCl is used as the tracer in the residence time distribution
experiments. The suitability of KCl as a tracer has been reported
earlier.[16,26] The tracer is injected at a point located 50 mm below the
top of the glass column. Care is taken to ensure that the entire volume of
the tracer is injected as quickly as possible. An online conductivity meter
(range of 0–2 S/cm) is located at the bottom of the active section (50 mm
above the base of the glass column) to continuously record the conductivity
of the continuous phase. The increase in conductivity above the back­
ground value is entirely due to the presence of the tracer. A data acquisition
system stores the data from the conductivity meter with a temporal resolu­
tion of 1 sec. Experiments are performed for different continuous and
dispersed phase flow rates and for different pulsing amplitudes. Pulsing
frequency is kept unchanged at 1 Hz. Pulsing duty cycle is kept 30%.
Physical properties of the phase system are reported elsewhere.[41]
6 N. SEN ET AL.

Computational model
The Euler–Euler method is used to model liquid-liquid two-phase flow in
PSPC. This method is widely used to simulate dispersed two-phase
flows.[35,36,42] The model essentially solves conservation equations of momen­
tum and mass of both phases and assumes the phases as interpenetrating
continua. The phase fraction of the dispersed phase in each computational cell
is calculated by solving a convection-diffusion transport equation for the phase
fraction itself. The interphase momentum exchange is considered in the
momentum equations of the two phases.
Interphase momentum exchange is represented in terms of interphase
exchange coefficient. The exchange of momentum between the phases is
only through the drag force which is quantified using a closure equation
(drag model). The drag model used in this work is of the form proposed by
Schiller and Naumann.[43]
Turbulence has been modelled by using the mixture k-ε model in which
the turbulence equations are solved for the mixture as a whole. This
approach reduces the number of equations to be solved, as turbulence
equations are not solved for each phase. One major limitation of most of
the previous studies on CFD simulations of PSPCs is the assumption of
monodispersed drops. With this assumption, the information on coales­
cence and redispersion of droplets, which are continuously taking place
inside the column, is lost. Thus, the local drop dynamics is not captured
with the assumption of monodispersed droplets. In our recent work, we
had addressed this issue and coupled population balance equations (PBE)
along with flow and turbulence equations to do away with the assumption
of monodispersed droplets.[39] Local drop size distribution in a liquid–
liquid dispersion depends on local velocity, breakage and coalescence
rates of droplets. Breakage and coalescence rates, in turn, depend on
physical properties of the phase system considered, local turbulent energy
dissipation rate and dispersed phase holdup. This physics is captured by the
population balance equations. A method of classes has been used and one
drop number density conservation equation has been solved for each class
of drops (class diameter L) in a computational cell. The pulsing action is
introduced into the computational model using an user-defined function.
The relevant governing equations are summarized in Table 1 .
Estimation of axial dispersion in continuous phase by using the CFD model
essentially involves solving an additional species transport equation in the
continuous phase as given by Eqn. (18).

@
ðαc CÞ þ Ñ:ðαc Uc CÞ ¼ Dc Ñ2 ðαc CÞ (18)
@t
Table 1. Governing equations.
@ Continuity equation (ith phase) (1)
@t ðαi ρi Þ þ Ñ:ðαi ρi Ui Þ ¼ 0
n
P Momentum conservation equation (ith phase) (2)
@
@t ðαi ρi Ui Þ þ Ñ:ðαi ρi Ui Ui Þ ¼ αi Ñp αi Ñ:τ i þ αi ρi g þ Rij
i¼1
T T Stress tensor term for Eqn. (2) (3)
τ i ¼ μi ðÑUi þ ÑUi Þ þ μt;m ðÑUi þ ÑUi Þ
P P
Rij ¼ �Kij ðUi �Uj Þ Interfacial drag force term for Eqn. (2) (4)
�! ! � Interphase momentum exchange coefficient for Eqn. (4) (5)
3αi αj ρj CD � U i U j �
Kij ¼ 4dp
� �
@ μt;m Conservation of turbulent kinetic energy of mixture k-ε model of turbulence (6)
@t ðρ m kÞ þ Ñ:ðρ m Um kÞ ¼ Ñ: σk Ñk þ Gk;m ρm ε
� � �
@ μt;m ε Conservation of turbulent energy dissipation rate of mixture k-ε model of turbulence (7)
@t ðρm εÞ þ Ñ:ðρm Um εÞ ¼ Ñ: σε Ñε þ k C1ε Gk;m C2ε ρm ε
Pn
αi ρi Ui
Expressions for mixture density, velocity and turbulent viscosity for use in Eqns. (3), (6–7) (8)
Pn 2
i¼1
ρm ¼ αi ρi ; Um ¼ P n ; μt;m ¼ ρm Cμ kε
i¼1 αi ρi
i¼1
� �
T Turbulent kinetic energy generation term for use in Eqns. (6,7) (9)
Gk;m ¼ μt;m ÑUm þ ÑUm : ÑUm
@

@t fn ðL; t Þ g þ Ñ: Ui :nðL; t Þ ¼ Ba ðL; tÞ Da ðL; tÞ þ Bb ðL; tÞ Db ðL; tÞ Conservation equation for drop population in each of L classes (10)
ð p3 ffiffi
L= 2 Birth rate of droplet of size L due to coalescence of smaller drops (11)
1=3
Ba ðL; tÞ ¼ βfðL3 λ3 Þ ; λgnfðL3 λ3 Þ1=3 ; tgnðλ; tÞdλ
0
ð1
Birth rate of droplet of size L due to breakage of larger drops (12)
Ba ðL; tÞ ¼ aðλÞbðLjλÞnðλ; tÞdλ
L
1 Death rate of droplet of size L due to coalescence with other drops (13)
Da ðL; tÞ ¼ nðL; tÞ ò βðL; λÞnðλ; tÞdλ
0
b Death rate of droplet of size L due to breakage into smaller drops (14)
D ðL; tÞ ¼ aðLÞnðL; tÞ n o
ε1=3 Þ2 Closure term (breakage kernel) for Eqns. (12) and (14) (Hsia and Tavlarides, 1980) (15)
ÞL2=3
aðLÞ ¼ C1 ð1þϕ exp C2 ρσðε1þϕ
2=3 L5=3
d

C1 ¼ 4:87 � 10 3 C2 ¼ 5:52 � 10 2
h i � � �4 � Closure term (coalescnce kernel) for Eqns. (11) and (13) from Hsia and Tavlarides, 1980 (16)
ε1=3 2 2=3 2=3 1=2 μc ρ c ε Lλ
βðL; λÞ ¼ C3 ð1þϕ Þ ðL þ λÞ ðL þ λ Þ exp C 4 σ2 ð1þϕÞ3 Lþλ

C3 ¼ 2:17X10 4
2
SOLVENT EXTRACTION AND ION EXCHANGE

C3 ¼ 2:28X1013 m
� 3 �2 � �2
L3 Closure term (daughter droplet distribution) for Eqn. (12) (Hsia and Tavlarides, 1980)[44] (17)
bðL; λÞ ¼ 30 Lλ3 1 λ3
7
8 N. SEN ET AL.

where C is the tracer concentration (in the continuous phase), Uc is the


velocity field of the continuous phase, αc is the continuous-phase volume
fraction and Dc is the effective diffusion coeffcient of the tracer in the con­
tinuous phase. Effective diffusion coefficient is the sum of molecular diffusion
coefficient (Dm) and eddy (turbulent) diffusion coefficient (Dt), as expressed by
Eq. (19).

1 μt;m
Dc ¼ Dm þ (19)
Sct ρm

The value of turbulent Schmidt number (Sct) is taken as 0.7. The second term
on the right hand side of Eq. (19) represents the eddy diffusion coefficient.
Initially, two-phase flow field is obtained by solving all governing equations
except the species transport equation. While these equations are solved,
average values of dispersed phase holdup and Sauter mean diameter in the
active section are monitored. Solution is continued till steady state values of
dispersed phase holdup and Sauter mean drop diameter are obtained.
Thereafter, a fully coupled unsteady simulation, in which species transport
equation is additionally solved, is carried out after applying a pulse tracer input
to obtain time-varying profile of tracer concentration at a location below the
bottom most plate in the computational domain. This essentially provides the
concentration curve (C-curve) which is normalized to obtain the E-curve. The
first moment (tm), the second moment (σ2) and dimensionless second moment
(σ 2θ Þ of the E-curve are evaluated by using Eq. 20.[45] The closed-closed
boundary condition is used in line with the published literature related to
pulsed columns.[13,46] The relationship between dimensionless second
moment and Peclet number is given by Eq. (21) which can be used to obtain
Peclet number once the value of σ 2θ is known from Eq. (20).
ð1 ð1
σ2
tm ¼ 2
tEðtÞdt; σ ¼ ðt tm Þ2 EðtÞdt; σ θ 2 ¼ (20)
0 0 tm 2

From σ θ 2 , Pe is estimated by solution of Eq. (21).[45]

2 2 �
σθ2 ¼ þ 2 1 e Pe
(21)
Pe Pe
From Pe, axial dispersion coefficient (Da) is estimated by using Eq. (22).

Pe
Da ¼ (22)
L:Vc
Da, Pe, L and Vc are axial dispersion coefficients of the continuous phase,
Peclet number, characteristic length and superficial flow velocity of the con­
tinuous phase, respectively.
SOLVENT EXTRACTION AND ION EXCHANGE 9

Figure 2. Meshed computational domain used for CFD modelling (left), mesh in two interplate
spaces (right).

In the present work, a 2D computational model with a reduced number of


plates (5 plates), shown in Figure 2, has been considered to limit the size of the
computational domain and the resulting computational time. Suitability of
using 2D model and reduced number of plates for CFD modelling of air pulsed
columns has been reported earlier.[26,28,32,34,37,39] Unsteady state simulations
are carried out with a time step size of 0.01 sec corresponding to Courant
number less than 0.5 in all cases. The computational domain used in the
present work is the same as in our previous work which describes the geometry
in detail.[37] For brevity, details are not being provided here. In this work, grid
density of 1.027 × 106 cells/m2 is used. This grid density is based on our
previous study on CFD modelling of single-phase flow in PSPC in which
detailed grid independence test was done.[28] Similar grid density has been
used in our recent work on two-phase flow in PSPC.[37]

Results and discussion


Experimental investigation of continuous phase axial dispersion

Axial dispersion in the continuous phase is considered in this work. The


temporal variation of the concentration of the tracer (KCl) at the outlet
gives the concentration curve (C-curve) which is transformed into the corre­
sponding E-curve by normalizing the concentration values by the total quan­
tity of the tracer injected. The details of normalization method are given
elsewhere.[45] Some preliminary experiments were carried out to understand
the effect of tracer concentration (0.5, 1.5 and 2.5 M) and injected tracer
volume (2.5, 5 and 7 mL) on the resultant E-curve. These tests indicated that
the E-curve becomes independent of the tracer concentration when the tracer
concentration is more than 1.5 M. Even though a concentration of 1.5 M was
sufficient, we used a KCl concentration of 2.5 M to increase the sensitivity of
10 N. SEN ET AL.

the conductivity data. It was also observed experimentally that the injected
volume of tracer should be lower than 5 mL. Thus, a volume of 2.5 mL was
used in the experiments. The normalized E-curve was then analyzed to obtain
the first and the second moments of the E-curve, which were then used to
evaluate Peclet number and axial dispersion coefficient (Da). Experiments
were carried out for different values of dispersed phase velocity
(0.0062–0.012 m/s), continuous phase velocity (0.0055–0.011 m/s) and pulsing
velocity (0.022 to 0.033 m/s). The pulsing frequency was kept constant at 1 Hz.
Figure 3a shows the experimentally obtained normalized E-curves for two
pulsing velocities (0.011 and 0.033 m/s). When pulsing velocity was varied, the
continuous phase velocity and dispersed phase velocity were maintained at
0.0055 m/s and 0.0062 m/s, respectively. As can be seen, at higher pulsing
velocity the spread of the E-curve is more and the peak value is less which
signify higher axial dispersion at higher pulsing velocity. The E-curves in both
the cases were further analyzed to obtain values of the axial dispersion
coefficient for the two cases. Axial dispersion coefficients for Af = 0.011 and
0.033 m/s are evaluated to be 0.000175 and 0.000336 m2/sec, respectively. This
clearly indicates that the axial mixing is more for higher pulsing velocity. One
important feature of the E-curve is that, it is not smooth. This occurs for two
reasons. First, the pulsatile nature of flow creates to and fro movement of the
tracer which is also sensed by the online conductivity sensor. Secondly, under
two-phase flow conditions, there would be organic drops that would come
now and then and hit the sensor tip. As organic phase is non-conducting, the

Figure 3. (a) Normalized E-curves for two different pulsing velocities (Vc = 0.0055 m/s, Vd
= 0.0062 m/s). Variation of continuous phase axial dispersion coefficient with (b) pulsing velocity
(Vc = 0.0055 m/s, Vd = 0.0062 m/s), (c) dispersed phase velocity (Af = 0.022 m/s, Vc = 0.0055 m/s),
and (d) continuous phase velocity (Af = 0.022 m/s, Vd = 0.0062 m/s).
SOLVENT EXTRACTION AND ION EXCHANGE 11

value of conductivity would momentarily reduce till the time the drop is
displaced. This would also cause fluctuation in conductivity values and in
the resultant E-curve.
Figure 3b shows the variation of axial dispersion coefficient with pulsing
velocity (Af). An increase in Af is clearly seen to increase continuous phase
axial dispersion coefficient. An increase in Af from 0.011 to 0.033 m/s (200%
increase) increased axial dispersion coefficient from 0.000175 to 0.000336 m2/
sec (91% increase). For the same values of Vc and Vd, an increase in Af would
increase to and fro axial movement of fluid which would in turn significantly
increase turbulence intensity and back-mixing in the column leading to an
increase in axial dispersion. It may also be mentioned here that increase of Af
from 0.011 m/s to 0.033 m/s also increased the mean residence time of the
continuous phase from 53.8 to 70.9 sec which can clearly be attributed to the
increased level of turbulence in the column. The greater the level of turbu­
lence, the greater is the degree of back mixing/recirculations in the column,
which would retain the tracer for a longer time in the column leading to larger
values of mean residence time for constant Vc and Vd.
Figure 3c shows the variation of axial dispersion coefficient with dispersed
phase velocity (Vd). When dispersed phase velocity was varied, the continuous
phase velocity and pulsing velocity were maintained at 0.0055 m/s and
0.022 m/s, respectively. It is observed that axial dispersion coefficient of the
continuous phase decreases with increase in dispersed phase velocity. As Vd is
raised from 0.0061 to 0.012 m/s (about 100% increase), axial dispersion
coefficient reduces from 0.00030 to 0.000275 m2/sec (about 11% reduction).
There are two effects at play when Vd is increased. An increase in Vd increases
dispersed phase holdup and thus decreases the space available for the contin­
uous phase to move down thereby increasing local velocities and hence
strength of recirculations in the continuous phase. This tends to increase
axial mixing in the continuous phase. But the presence of more of dispersed
phase drops tends to break a large recirculation of the continuous phase into
smaller multiple ones. This effect tends to reduce axial dispersion/mixing in
the continuous phase. A combination of these two factors governs the nature
of dependence of axial mixing on dispersed phase velocity. In PSPC, due to the
presence of holes in the plates, drops of the dispersed are injected into the
recirculations of the continuous phase. This breaks the recirculations in the
continuous phase into smaller localized recirculations. Due to this effect the
axial dispersion in the continuous phase reduces with an increase in Vd. It may
be mentioned here that an increase in Vd from 0.006 to 0.012 m/s also reduces
the mean residence time of the continuous phase from 65 to 55 sec. This is
attributed to the fact that with increase in Vd the dispersed phase holdup in the
column increases which reduces the space available for the continuous phase
thereby leading to faster exit of the continuous phase from the column
resulting in smaller mean residence time.
12 N. SEN ET AL.

Figure 3d shows the variation of axial dispersion coefficient of the contin­


uous phase with continuous phase velocity. Dispersed phase velocity and
pulsing velocity were maintained at 0.0062 m/s and 0.022 m/s, respectively,
in these experiments. It is observed that axial dispersion coefficient keeps on
increasing with increase in Vc. As Vc is raised from 0.0055 to 0.011 m/s (100%
increase), axial dispersion coefficient increases from 0.00030 to 0.00066 m2/s
(116% increase). This is because of the fact that as Vc is increased, the
recirculations in the continuous phase become stronger causing enhancement
in axial mixing. Increase in continuous phase velocity from 0.0055 to 0.011 m/
s also reduces mean residence time of the continuous phase from 65 to 51 sec.
As the Vc is raised, the time spent by the continuous phase in the column
reduces leading to reduction in mean residence time with increase in contin­
uous phase velocity. Going by the percentage change figures, among the
operating variables, continuous phase velocity affects the continuous phase
axial dispersion the most and dispersed phase velocity affects the continuous
phase axial dispersion the least.

Validation of the model


The CFD model is used to obtain the periodic steady state values of dispersed
phase holdup and Sauter mean drop diameter for a given set of operating
parameters. The two-phase flow field is established before virtual tracer study
is initiated. This methodology is similar to what is adopted in experiments also
in which the tracer is injected after the two-phase flow is properly established.
Figure 4a shows the comparison of the normalized E-curve obtained
experimentally and that predicted by the CFD model. Values of Vc, Vd and
Af for the case shown are 0.0055, 0.0062 and 0.022 m/s, respectively. Some
superimposed fluctuations could be observed on the experimental data. These
were due to the reasons mentioned earlier namely pulsing induced to and fro
movement of the tracer and momentary obstruction of the sensor tip by drops
of the organic phase. However, the congruence between the predicted and
experimental data is reasonably good. After analyzing the respective E-curves,
values of continuous phase axial dispersion coefficient for the experimental
E-curve and predicted E-curve are estimated to be 0.00030 and 0.00031 m2/s,
respectively, for the case shown in Figure 4a.
Detailed validation of the model was carried out for different values of Af,
Vd and Vc. Figure 4b shows the comparison of predicted values of continuous
phase axial dispersion coefficient with experimental values for different pul­
sing velocities. Values of dispersed and continuous phase velocities were fixed
at 0.0062 m/s and 0.0055 m/s, respectively. Figure 4c shows the comparison of
continuous phase axial dispersion coefficient predicted from the model and
corresponding experimental values for different dispersed phase velocities.
Values of continuous phase velocity and pulsing velocity were fixed at
SOLVENT EXTRACTION AND ION EXCHANGE 13

Figure 4. (a) Comparison of experimental and predicted normalized E-curves (Af = 0.022 m/s, Vc
= 0.0055 m/s, Vd = 0.0062 m/s). Comparison of predicted and experimental values of continuous
phase axial dispersion coefficient for (b) different pulsing velocities (Vc = 0.0055 m/s, Vd
= 0.0062 m/s), (c) different dispersed phase velocities (Af = 0.022 m/s, Vc = 0.0055 m/s), and (d)
different continuous phase velocities (Af = 0.022 m/s, Vd = 0.0062 m/s). (e) Parity plot of predicted
and experimentally obtained values of continuous phase axial dispersion coefficients.

0.0055 m/s and 0.022 m/s, respectively. Similarly, Figure 4d shows the com­
parison of axial dispersion coefficient predicted from the model and corre­
sponding experimental values for different continuous phase velocities. Values
of dispersed phase velocity and pulsing velocity were fixed at 0.0062 m/s and
0.022 m/s, respectively. As evident from Figure 4b-d, the predictions of axial
dispersion coefficient by the CFD model are close to the experimental values.
Quantitatively, it is found that the average absolute relative error in prediction
of continuous phase axial dispersion coefficient is 2.56%. This is also shown in
Figure 4e, which presents the parity plot between the experimental and model-
predicted values of continuous phase axial dispersion coefficient. It is observed
that all the points are well within ±10% error bands marked in the figure as
dotted lines.
14 N. SEN ET AL.

Figure 5. Spatial variation of dispersed phase holdup inside the column at the postive peak and
the negative peak of a pulsing cycle (Af = 0.022 m/s; Vd = 0.0062 m/s, Vc = 0.0055 m/s).

Local two-phase flow hydrodynamics


In this section, we use the validated CFD model to obtain insights into the
local hydrodynamics inside the column. Figure 5 shows the spatial variation of
dispersed phase holdup at two different instants of time during a pulsing cycle.
The figure on the left side is at the instant of positive peak of the pulse while
the figure on the right side is at the instant of the negative peak of the pulse. It
is interesting to note that within a pulsing cycle local distribution of the
dispersed phase changes significantly. It can be clearly observed that during
the positive peak of the pulsing cycle the dispersed phase oozes upwards
through the sieve holes while during the negative peak the dispersed phase
accumulates below the plates. This is expected as the overall movement of the
phases is synchronized with pulsing. The net flow of the dispersed phase
(lighter phase) is from the bottom to the top. The upstroke of the pulse
facilitates this net upward flow of the dispersed phase and as a result the
dispersed phase is seen to be oozing out of the holes upward at the positive
peak of the pulse. The net flow of the continuous phase (heavier phase) is from
the top to the bottom. This net flow of the continuous phase is facilitated by
the downstroke of the pulse. As a result, the continuous phase moves through
the holes downward at the negative peak of the pulse while the dispersed phase
tends to accumulate below the plate.
Figure 6 shows the spatial variation of Sauter mean drop diameter, and
turbulent energy dissipation rate in an interplate space (between 4th and 5th
plate from bottom). The instant of time corresponds to the positive peak of the
SOLVENT EXTRACTION AND ION EXCHANGE 15

Figure 6. Spatial variation of Sauter mean drop diameter and turbulence dissipation rate at the
positive peak of the pulsing cycle (Af = 0.022 m/s; Vd = 0.0062 m/s, Vc = 0.0055 m/s).

pulsing cycle. The drop diameter is smaller in the vicinity of the holes. These
locations are the ones where local values of turbulent energy dissipation rate
are high. High values of turbulence energy dissipation rates lead to higher rates
of drop breakage leading to generation of smaller drops in regions around the
sieve holes. As the drops move up and away from the holes they tend to
coalesce leading to formation of larger drops.
Figure 7 shows the comparison of dispersed phase holdup inside the
column. The comparison is done for two different dispersed phase velocities.
The profiles are at the positive peak of the pulse. An increase in dispersed
phase holdup is clearly observed for higher dispersed phase velocity. In fact,
for a dispersed phase flow velocity of 0.0075 m/s a thin layer of dispersed phase
is clearly seen to accumulate below the plates. The dispersed phase is seen to
ooze out of the sieve holes which is expected at the positive peak of the pulsing
cycle.
Figure 8 shows the comparison of the spatial variation of Sauter mean
diameter at two different values of dispersed phase velocity (i.e. 0.0062 m/s
and 0.0075 m/s). The comparison is done at the positive peak of the pulsing
cycle. A clear increase in Sauter mean drop diameter is observed as the
dispersed phase velocity is raised. As dispersed phase velocity is increased
more and more drops are packed below the sieve plate which leads to an
increase in dispersed phase hold up below the plate (Figure 7). This in turn
leads to a higher coalescence rate leading to larger drops.[39] Furthermore, the
increase in Sauter mean drop as the drops move away from the holes in the
bottom sieve plate is clearly revealed in both the cases, though the difference is
more prominent for the case of higher dispersed phase velocity.
16 N. SEN ET AL.

Figure 7. Dispersed phase holdup profiles for two different values of dispersed phase velocity at
the positive peak of pulsing cycle (Vc = 0.0055 m/s; Af = 0.022 m/s).

Figure 9 shows the comparison of dispersed phase holdup and Sauter mean
drop diameter in the interplate spacing between 4th and 5th plates for two
different values of pulsing velocity. The profiles are shown at the positive peak
of the pulsing cycle. Dispersed phase hold up is clearly seen to increase with
increase in pulsing velocity whereas Sauter mean drop diameter decreases with
an increase in pulsing velocity. A reduction is drop size at higher pulsing
velocity is attributed to the increased drop breakage rates due to higher local
values of trubulent energy dissipation rate. The fact that local turbulent energy
dissipation rates are higher at a higher value of pulsing velocity is clearly seen
in Figure 10 which compares spatial variation of turbulent energy dissipation
rate for two pulsing velocities (0.033 and 0.022 m/s). A reduction in drop size
leads to a larger dispersed phase hold up as smaller drops have a greater
tendency to get trapped in the recirculations present in the continuous phase.
The CFD-PB predictions of dispersed phase hold up and Sauter mean
drop diameter for three different values of Af and three different values of
Vd are listed in Table 2. The values predicted by CFD-PB (averaged over
a time cycle) are compared against those predicted by the correlations
SOLVENT EXTRACTION AND ION EXCHANGE 17

Figure 8. Sapatial variation of Sauter mean drop diameter in an inter-plate space for two different
values of dispersed phase velocity at the positive peak of pulsing cycle (Vc = 0.0055 m/s;
Af = 0.022 m/s).

Figure 9. Spatial variation of dispersed phase hold up (top) and Sauter mean drop diameter
(bottom) in an inter-plate space for two different values of pulsing velocity at positive peak of
pulsing cycle (Vc = 0.0055 m/s; Vd = 0.0062 m/s).
18 N. SEN ET AL.

Figure 10. Sapatial variation of turbulent dissipation rate in an inter-plate space for two different
values of pulsing velocity at positive peak of pulsing cycle (Vc = 0.0055 m/s; Vd = 0.0062 m/s).

Table 2. Comparison of CFD-PB predicted dispersed phase holdup and Sauter mean drop diameter
vis-a-vis those predicted from correlations.
Dispersed phase holdup (%) Sauter mean drop diameter (mm)
Parameter CFD-PB Correlation of Eq. (23) CFD-PB Correlation of Eq. (24)
Af (m/s) 0.022 12.4 13.31 1.38 1.63
(Vc = 0.0055 m/s; 0.0278 21.7 16.72 0.93 1.35
Vd = 0.0062 m/s) 0.033 24.1 20.52 0.92 1.18
Vd (m/s) 0.0062 12.4 13.31 1.38 1.63
(Af = 0.022 m/s; 0.0075 21.2 16.16 1.61 1.63
Vc = 0.0062 m/s) 0.0092 26.3 20.01 1.75 1.63

available in literature for dispersed phase hold up[41] and Sauter mean drop
diameter.[47] The above correlations were chosen as the liquid-liuqid phase
system and the dimension of the column used in the present work are very
similar/same as used in the studies in which these correlations are reported.
The correlations are given by Eqs. 23 and 24.
Results of Table 2 clearly show that CFD-PB predicted values of holdup and
Sauter mean drop diameter match reasonably well with those predicted by the
correlations for the same. Average absolute relative errors in prediction of
dispersed phase holdup and Sauter mean drop are 20.68% and 19.90%.
However, validation of the CFD-PB results with actual experimental data
was not done as experimental data on dispersed phase holdup and Sauter
mean drop diameter were not recorded during experimental runs in the
present work.
� � ��
ϕ ¼ 36:5exp 39:35�Af ðAf Þm � Vd1:02 Vc0:02 Δρ 0:23 0:52
μd dh 0:3 e 0:4 h 0:4
SOLVENT EXTRACTION AND ION EXCHANGE 19

!0:33
3 σΔρ1=4 e
ðAf Þm ¼ 9:69 � 10 3=4
(23)
μd

� �0:4
σ 0:8
d32 ¼ 0:21 e0:48 dh0:26 h0:34 ðAf Þ (24)
ρc

Figure 11 shows the velocity vectors of the continuous phase at two


different dispersed phase velocities (0.0062 m/s and 0.0075 m/s). The
comparison is done at the positive peak of the pulsing cycle. Values of
continuous phase velocity and pulsing velocity were maintained constant at
0.0055 m/s and 0.022 m/s, respectively. At low values of dispersed phase
velocity, two well-defined counter-rotating recirculatory loops are formed in
the continuous phase. However, as dispersed phase velocity is raised, these
loops are no longer well defined and are broken down into several smaller
ones. For the same continuous phase velocity and pulsing velocity, the
strength of the recirculations is also reduced when dispersed phase velocity
is increased, as evident qualitatively by lower local velocities in the re
circulatory loops for higher dispersed phase velocity. Such a difference
can be attributed to the fact that at high values of dispersed phase velocity
a larger extent of dispersed phase will be pushed into the continuous phase
streams thereby breaking up the continuous phase recirculations. Hence,
a large, sustained recirculation in the continuous phase is effectively broken
down into smaller and weaker ones as dispersed phase velocity is increased.
This is also evident from Figure 12 which compares the dispersed phase
velocity vectors for two different dispersed phase velocities. It is clearly seen

Figure 11. Continuous phase velocity vectors for two different values of dispersed phase velocity
at positive peak of pulsing cycle (Vc = 0.0055 m/s; Af = 0.022 m/s).
20 N. SEN ET AL.

Figure 12. Dispersed phase velocity vectors for two different values of dispersed phase velocity at
the positive peak of pulsing cycle (Vc = 0.0055 m/s; Af = 0.022 m/s).

that at low dispersed phase velocity the dispersed phase preferentially


moves up through the centre of the column while significant re
circulations are set up in continuous phase. Thus, the flow pattern of the
dispersed phase is not uniform across the cross-section.
As the dispersed phase velocity is increased from 0.0062 m/s to
0.0075 m/s, the dispersed phase starts moving much more uniformly
across the cross-section of the column. As the dispersed phase is now
being pushed out from the holes more uniformly across the column, it
tends to break the sustained recirculations in the continuous phase as
observed in Figure 11 for the case of higher dispersed phase velocity.
This decrease in size and strength of recirculations in continuous phase
essentially leads to reduced axial mixing in the column and can be interpreted
as the reason for reduction in axial dispersion coefficient with increase in
dispersed phase velocity.

Conclusions
A Computational Fluid Dynamics model is used to predict axial dispersion in
continuous phase in liquid-liquid two-phase flow in a pulsed sieve plate
column (PSPC). Drop size and dispersed phase holdup are obtained by
coupling population balance model with two-phase CFD model. Species
transport equation is solved along with the equations of CFD and population
balance models to computationally predict residence time distribution curve
which is analyzed to obtain axial dispersion coefficient of the continuous
phase. The model is validated with the experimental data generated by
SOLVENT EXTRACTION AND ION EXCHANGE 21

conducting experiments on a 3-inch diameter PSPC to obtain axial dispersion


coefficient in continuous phase for different values of continuous phase,
dispersed phase and pulsing velocities. The average absolute relative error in
prediction of axial dispersion coefficient is found to be 2.55%. Axial dispersion
coefficient is found to increase with increase in pulsing velocity and contin­
uous phase velocity while it is found to decrease slightly with increase in
dispersed phase velocity.
The validated model is used to gain detailed insights into the hydrody­
namics prevailing inside the column. Smaller drops are observed in the
regions near the sieve holes which are characterised by higher values of
turbulent energy dissipation rate. Dispersed phase holdup is found to
increase with increase in pulsing velocity and dispersed phase velocity.
Sauter mean drop diameter reduces with an increase in pulsing velocity
due to an increase in turbulent energy dissipation rate at higher pulsing
velocity. Values of dispersed phase holdup and Sauter mean drop diameter
predicted by CFD-PB model were shown to be reasonably close to those
obtained by using previously reported correlations. At low values of dis­
persed phase velocity sustained recirculations in the continuous phase are
observed while the dispersed phase preferentially moves through the centre
of the column. However, as the dispersed phase velocity is raised, the re
circulations in the continuous phase reduce in the span as well as strength
and the dispersed phase moves more uniformly across the column cross-
section. This observation explains the reason for reduction in axial disper­
sion coefficient with increase in dispersed phase velocity as was observed in
the experiments as well as in numerical simulations.

Nomenclature
A Pulsing amplitude [L]
C Tracer concentration [M/L3]
C1ε, C2ε, Cµ Constants in standard k-ε model [-]
C1 Constant in drop breakage kernel [-]
C2 Constant in drop breakage kernel [-]
C3 Constant in aggregation (coalescence) kernel [-]
C4 Constant in aggregation (coalescence) kernel [1/L2]
CD Drag coefficient [-]
Dc Effective diffusivity of tracer [L2/T]
Da Axial dispersion coefficient [L2/T]
dh Hole diameter [L]
E(t) Exit age distribution function [1/T]
E(θ) Dimensionless exit age distribution function [-]
e Fractional open area of sieve plate [-]
f Pulse frequency [1/T]
Gk,m Kinetic energy generation term for the mixed phase [1/LT3]
h Interplate spacing [L]
22 N. SEN ET AL.

k Turbulent kinetic energy [L2/T2]


Kij Interphase momentum exchange coefficient [M/L3T]
Rij Interphase exchange force [M /L2T2]
n(L,t) number of drops of size L per unit volume at time t [1/L3]
p Static pressure [M/LT2]
Pe Peclet number (continuous phase) [-]
tm Mean residence time [T]
Ui Velocity vector of ith phase [L/T]
Um Mixture velocity vector [L/T]
Vd Dispersed phase superficial flow velocity [L/T]
Vc Continuous phase superficial flow velocity [L/T]

Greek letters
αi Phase fraction of ith phase [-]
ε Turbulent energy dissipation rate [L2/T3]
ρ Density [M/L3]
ϕ Dispersed phase holdup [-]
µ Viscosity [M/LT]
σ Interfacial tension [M/T2]
σ2 Second moment of E curve [T2]
σθ2 Dimensionless second moment of E curve [-]

Subscript
c Continuous phase
d Dispersed phase
m Mixture

ORCID
K.K. Singh http://orcid.org/0000-0002-5155-0462
A. W. Patwardhan http://orcid.org/0000-0001-9893-1249

References
[1] Griffith, W. L.; Jasny, G. R.; Tupper, H. T. The Extraction of Cobalt from Nickel in a Pulse
Column; Report No. K-972, Massachusetts Institute of Technology (1952).
[2] Golding, J. A.; Lee, J. Recovery and Separation of Cobalt and Nickel in a Pulsed
Sieve-plate Extraction Column. Ind. Eng. Chem. Process Des. Dev. 1981, 20(2),
256–261. DOI: 10.1021/i200013a013.
[3] Ferreira, A. E.; Agarwal, S.; Machado, R. M.; Gameiro, M. L. F.; Santos, S. M. C.; Reis, M.
T. A.; Ismael, M. R. C.; Correia, M. J. N.; Carvalho, J. M. R. Extraction of Copper from
Acidic Leach Solution with Acorga M5640 Using a Pulsed Sieve Plate Column.
Hydrometallurgy. 2010, 104(1), 66–75.
SOLVENT EXTRACTION AND ION EXCHANGE 23

[4] Gameiro, F. M. L.; Machado, R. M.; Ismael, M. C.; Reis, M. A.; Carvalho, J. M. Copper
Extraction from Ammoniacal Medium in a Pulsed Sieve-plate Column with LIX 84-I.
J. Hazard. Mater. 2010, 183(1), 165–175. DOI: 10.1016/j.jhazmat.2010.07.006.
[5] Ban, Y.; Hotoku, S.; Tsubata, Y.; Morita, Y. Recovery of U and Pu from Nitric Acid Using
N,N-di(2-ethylhexyl)butanamide (DEHBA) in Mixer-settler Extractors. Solvent Extr. Ion
Exch. 2013, 31(6), 590–603. DOI: 10.1080/07366299.2013.806744.
[6] Chen, H.; Taylor, R.; Jobson, M.; Woodhead, D.; Masters, A. Development and
Validation of a Flowsheet Simulation Model for Neptunium Extraction in an
Advanced PUREX Process. Solvent Extr. Ion Exch. 2016, 34(4), 297–321. DOI:
10.1080/07366299.2016.1185853.
[7] Farakte, R. A.; Hendre, N. V.; Patwardhan, A. W. CFD Simulations of Two-phase Flow
in Asymmetric Rotary Agitated Columns. Ind. Eng. Chem. Res. 2018, 57(50),
17192–17208. DOI: 10.1021/acs.iecr.8b02720.
[8] Weber, B.; Schneider, M.; Görtz, J.; Jupke, A. Compartment Model for Liquid-liquid
Extrcation Columns. Solvent Extr. Ion Exch. 2019, 38(1), 66–87. DOI: 10.1080/
07366299.2019.1691137.
[9] Chaturabul, S.; Wannachod, P.; Rojanasiraprapa, B.; Summakasipong, S.;
Lothongkum, A. W.; Pancharoen, U. Arsenic Removal from Natural Gas Condensate
Using a Pulsed Sieve Plate Column and Mass Transfer Efficiency. Sep. Sci. Technol. 2012,
47(3), 432–439. DOI: 10.1080/01496395.2011.614833.
[10] Liebermann, E.; Jealous, A. C. Application of the Pulse Column for Uranium Recovery
from K-25 Wastes; Oak Ridge National Laboratory, ORNL-1543, 1953.
[11] Schön, J.; Schmieder, H.; Kanellakopulos., B. Operating Experience with Minka: U-Pu
Separation in the Presence of Technetium with an Organic Continuous Pulsed Column.
Sep. Sci. Technol. 1990, 25(13–15), 1737–1750. DOI: 10.1080/01496399008050420.
[12] Yadav, R. L.; Patwardhan, A. W. Design Aspects of Pulsed Sieve Plate Columns. Chem.
Eng. J. 2008, 138(1–3), 389–415. DOI: 10.1016/j.cej.2007.06.015.
[13] Gonda, K.; Matsuda, T. Solvent Extraction Calculation Model for PUREX Process in
Pulsed Sieve Plate Column. J. Nucl. Sci. Technol. 1986, 23(10), 883–895. DOI: 10.1080/
18811248.1986.9735072.
[14] Nabli, M. A.; Guiraud, P.; Gourdon, C. Numerical Experimentation: A Tool to Calculate
the Axial Dispersion Coefficient in Discs and Doughnuts Pulsed Solvent Extraction
Columns. Chem. Eng. Sci. 1997, 52(14), 2353–2368. DOI: 10.1016/S0009-2509(96)00517-9.
[15] Li, N. N.; Ziegler, E. N. Effect of Axial Mixing on Mass Transfer in Extraction Columns.
Ind. Eng. Chem. 1967, 59(3), pp.30–36. DOI: 10.1021/ie51402a008.
[16] Rao, K. K.; Jeelani, S. A. K.; Balasubramanian, G. R. Backmixing in Pulsed Perforated
Plate Columns. Can. J. Chem. Eng. 1978, 56(1), pp.120–123. DOI: 10.1002/
cjce.5450560117.
[17] Jahya, A. B.; Pratt, H. C.; Stevens, G. W. Comparison of the Performance of a Plused Disc
and Doughnut Column with a Pulsed Sieve Plate Liquid Extraction Column. Solvent
Extr. Ion Exch. 2005, 23, 307–317. DOI: 10.1081/SEI-200045257.
[18] Prvcic, L. M.; Pratt, H. C.; Stevens, G. W. Axial Dispersion in Pulsed-, Perforated-plate
Extraction Columns. AIChE J. 1989, 35(11), pp.1845–1855.
[19] Baird, M. H. I.;. Axial Dispersion in a Pulsed Plate Column. Can. J. Chem. Eng. 1974, 52
(6), pp.750–757. DOI: 10.1002/cjce.5450520608.
[20] Yu, Y. H.; Kim, S. D. Axial Dispersion, Holdup and Flooding Characteristics in Pulsed
Extraction Columns. Can. J. Chem. Eng. 1987, 65(5), pp.752–758. DOI: 10.1002/
cjce.5450650507.
[21] Niebuhr, D.; Vogelpohl, A. Axial Mixing in Pulsed Sieve-plate Extraction Columns. Ger.
Chem. Eng. 1980, 3, p.264.
24 N. SEN ET AL.

[22] Noh, J. S.; Kim, S. D. Axial Dispersion in a Pulsed Extraction Column. Korean Chem.
Eng. Res. 1980, 18(3), pp.141-141.
[23] Tung, L. S.; Luecke, R. H. Mass Transfer and Drop Sizes in Pulsed-plate Extraction
Columns. Ind. Eng. Chem. Process Des. Dev. 1986, 25(3), pp.664–673. DOI: 10.1021/
i200034a012.
[24] Srinikethan, G.; Prabhakar, A.; Varma, Y. B. G. Axial Dispersion in Plate-pulsed
Columns. Bioprocess Eng. 1987, 2(4), pp.161–168. DOI: 10.1007/BF00387323.
[25] Kumar, A.; Hartland, S. Prediction of Continuous-phase Axial Mixing Coefficients in
Pulsed Perforated-plate Extraction Columns. Ind. Eng. Chem. Res. 1989, 28(10), pp.­
1507–1513. DOI: 10.1021/ie00094a013.
[26] Kolhe, N. S.; Mirage, Y. H.; Patwardhan, A. V.; Rathod, V. K.; Pandey, N. K.;
Mudali, U. K.; Natarajan, R. CFD and Experimental Studies of Single Phase Axial
Dispersion Coefficient in Pulsed Sieve Plate Column. Chemical Engineering Research
and Design. 2011, 89(10), 1909–1918.
[27] Xiaojin, T.; Guangsheng, L. CFD Simulations of Flow Characteristics in Pulsed-sieve-plate
Extraction Columns. Ind. Eng. Chem. Res. 2011, 50(2), 1110–1114. DOI: 10.1021/ie101788b.
[28] Sen, N.; Singh, K. K.; Patwardhan, A. W.; Mukhopadhyay, S.; Shenoy, K. T. CFD
Simulations of Pulsed Sieve Plate Column: Axial Dispersion in Single-phase Flow. Sep.
Sci. Technol. 2015, 50(16), 2485–2495.
[29] Retieb, S.; Guiraud, P.; Angelov, G.; Gourdon, C. Hold-up within Two-phase
Countercurrent Pulsed Columns via Eulerian Simulations. Chem. Eng. Sci. 2007, 62
(17), 4558–4572. DOI: 10.1016/j.ces.2007.04.043.
[30] Saini, R. K.; Bose, M. Stage Holdup of Dispersed Phase in Disc & Doughnut Pulsed
Column. Energy Procedia. 2014, 54, 796–803. DOI: 10.1016/j.egypro.2014.07.323.
[31] Yi, H.; Wang, Y.; Smith, K. H.; Fei, W. Y.; Stevens, G. W. CFD Simulation of
Liquid-liquid Two-phase Hydrodynamics and Axial Dispersion Analysis for a
Non-pulsed Disc and Doughnut Solvent Extrcation Column. Solvent Extr. Ion Exch.
2016, 34(6), pp.535–548. DOI: 10.1080/07366299.2016.1226025.
[32] Sarkar, S.; Singh, K. K.; Shenoy, K. T. Two-phase CFD Modeling of Pulsed Disc and
Doughnut Column: Prediction of Dispersed Phase Holdup. Sep. Purif. Technol. 2019a,
209, 608–622. DOI: 10.1016/j.seppur.2018.07.020.
[33] Sarkar, S.; Singh, K. K.; Shenoy, K. T. CFD Modeling of Pulsed Disc and Doughnut
Column: Prediction of Axial Dispersion in Pulsatile Liquid–Liquid Two-Phase Flow.
Ind. Eng. Chem. Res. 2019b, 58(33), pp.15307–15320. DOI: 10.1021/acs.iecr.9b01465.
[34] Sarkar, S.; Singh, K. K.; Shenoy, K. T. CFD-PB Modelling of Liquid–liquid Two-phase
Flow in Pulsed Disc and Doughnut Column. Solvent Extr. Ion Exch. 2020. DOI: 10.1080/
07366299.2020.1767360.
[35] Yadav, R. L.; Patwardhan, A. W. CFD Modeling of Sieve and Pulsed-sieve Plate
Extraction Columns. Chemical Engineering Research and Design. 2009, 87(1), 25–35.
DOI: 10.1016/j.cherd.2008.06.010.
[36] Din, G. U.; Chughtai, I. R.; Inayat, M. H.; Khan, I. H.; Qazi, N. K. Modeling of a Two-
phase Countercurrent Pulsed Sieve Plate Extraction Column—a Hybrid CFD and
Radiotracer RTD Analysis Approach. Sep. Purif. Technol. 2010, 73(2), 302–309. DOI:
10.1016/j.seppur.2010.04.017.
[37] Sen, N.; Singh, K. K.; Patwardhan, A. W.; Mukhopadhyay, S.; Shenoy, K. T. CFD
Simulation of Two-phase Flow in Pulsed Sieve-plate column–Identification of
a Suitable Drag Model to Predict Dispersed Phase Hold Up. Sep. Sci. Technol. 2016, 51
(17), 2790–2803. DOI: 10.1080/01496395.2016.1218895.
SOLVENT EXTRACTION AND ION EXCHANGE 25

[38] Sen, N.; Singh, K. K.; Patwardhan, A. W.; Mukhopadhyay, S.; Shenoy, K. T. CFD
Simulations to Predict Dispersed Phase Holdup in a Pulsed Sieve Plate Column. Sep.
Sci. Technol. 2018, 53(16), 2587–2600.
[39] Sen, N.; Singh, K. K.; Patwardhan, A. W.; Mukhopadhyay, S.; Shenoy, K. T. CFD-PB
Simulations of a Pulsed Sieve Plate Column. Prog. Nucl. Energy. 2019, 111, 125–137.
DOI: 10.1016/j.pnucene.2018.10.012.
[40] Yi, H.; Smith, K. H.; Fei, W. Y.; Stevens, G. W. CFD Simulation of Two-phase Flow in
Hybrid Pulsed Sieve-plate Solvent Extrcation Column: Prediction of Holdup and Axial
Dispersion Coefficients. Solvent Extr. Ion Exch. 2020, 38(1), pp.88–102. DOI: 10.1080/
07366299.2019.1691300.
[41] Sarkar, S.; Sen, N.; Singh, K. K.; Mukhopadhyay, S.; Shenoy, K. T. Effect of Operating and
Geometric Parameters on Dispersed Phase Holdup in Pulsed Disc and Doughnut and
Pulsed Sieve Plate Columns: A Comparative Study. Chem. Eng. Process. 2017, 118,
131–142. DOI: 10.1016/j.cep.2017.04.016.
[42] Wang, H.; Jia, X.; Wang, X.; Zhou, Z.; Wen, J.; Zhang, J. CFD Modeling of
Hydrodynamic Characteristics of a Gas–liquid Two-phase Stirred Tank. Appl. Math.
Modell. 2014, 38(1), 63–92.
[43] Schiller, L.; Naumann, Z. Vereines Deutscher Ingenieure. 1935, 77, 318.
[44] Hsia, M. A.; Tavlarides, L. L. A Simulation Model for Homogeneous Dispersions in
Stirred Tanks. Chem. Eng. J. 1980, 20(3), 225–236. DOI: 10.1016/0300-9467(80)80007-4.
[45] Fogler, H. S.; Elements of Chemical Reaction Engineering; Prentice Hall: New Delhi,
India, 1986.
[46] Torab-Mostaedi, M.; Safdari, J. Prediction of Mass Transfer Coefficients in a Pulsed
Packed Extraction Column Using Effective Diffusivity. Braz. J. Chem. Eng. 2009, 26(4),
pp.685–694. DOI: 10.1590/S0104-66322009000400007.
[47] Sarkar, S.; Sen, N.; Singh, K. K.; Mukhopadhyay, S.; Shenoy, K. T. Liquid-liquid
Dispersion in Pulsed Disc and Doughnut Column and Pulsed Sieve Plate Column:
A Comparative Study. Prog. Nucl. Energy. 2019c, 116, 76–86. DOI: 10.1016/j.
pnucene.2019.03.037.

You might also like