You are on page 1of 11

Chemical Engineering Science 63 (2008) 721 731

www.elsevier.com/locate/ces

Effect of ejector conguration on hydrodynamic characteristics


of gasliquid ejectors
S. Balamurugan, V.G. Gaikar, A.W. Patwardhan
Chemical Engineering Department, Institute of Chemical Technology, University of Mumbai, Matunga, Mumbai 400 019, India
Received 15 February 2006; received in revised form 25 September 2007; accepted 1 October 2007
Available online 9 October 2007

Abstract
Ejectors are gasliquid contactors that are reported to provide higher mass transfer rates than conventional contactors. Detailed experiments
were performed and computational uid dynamics (CFD) modeling studies were undertaken to understand the hydrodynamic characteristics
of the ejector geometry. The CFD model provides a basis for quantifying the effects of operating conditions on the ejector performance. CFD
studies shows that there is an optimum ratio of nozzle area to throat area (area ratio), at which the liquid entrainment rate is the highest. This
can lead to substantial economic benet in the industrial practice. The liquid entrainment rate correlates with pressure difference between the
water surface in the suction chamber and the throat exit for a wide variety of ejector geometries and operating conditions.
2007 Elsevier Ltd. All rights reserved.
Keywords: Hydrodynamics; Multiphase ow; Ejector; Entrainment; Pressure drop; CFD

1. Introduction
An ejector is a device in which a uid is pumped through a
nozzle at a high velocity (called the motive uid). The high velocity jet transports momentum to the outside uid and causes a
suction of the surrounding uid (called the entrained uid). The
mixing of the motive uid jet emerging from the nozzle and the
entrained uid in the mixing tube leads to the dispersion of one
phase into another in the throat of ejector. The diffuser section
after the mixing tube/throat helps in pressure recovery. Based
on the uids involved, three types of ejectors have been developed and been used over the years. The gasgas type ejectors are
used for the generation of vacuum. The liquidliquid ejectors
are used as mixing units in the mixersettler setup (Mukherjee
et al., 1988). Gasliquid ejectors use the jet of either gas or
liquid from the nozzle to entrain the other uid surrounding
the nozzle. The gasliquid ejectors are used in chemical industries for absorption and stripping (Ben Brahim et al., 1984) and
in biochemical industry. Ejectors produce high mass transfer

Corresponding author. Tel.: +91 22 2414 5616; fax: +91 22 2414 5614.

E-mail address: awp@udct.org (A.W. Patwardhan).


0009-2509/$ - see front matter 2007 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ces.2007.10.002

rates by generating small bubbles/droplets which can then be


injected into a reaction vessel thereby improving the contact
between phases (Cramers and Beenackers, 2001). As compared
to other gasliquid contacting systems, like stirred tanks and
bubble columns, ejectors have two distinct features. They (i)
entrain one uid without an external pumping agent and (ii)
produce higher values of volumetric mass transfer coefcient
because of high shear rates generated by the jet (Zahradnik
et al., 1997; Havelka et al., 2000).
In the chemical exchange process producing heavy water, a
synthesis gas mixture of nitrogen and hydrogen is contacted
with liquid ammonia at high pressure and low temperature
conditions. The deuterium extraction from the gas mixture into
the liquid ammonia takes place in the presence of a KNH2 as
a catalyst. Deuterium is present in gaseous hydrogen as HD at
a concentration of about 100 ppm. HD dissolves into the liquid
phase and reacts with ammonia to form deuteriated ammonia.
The rate of this exchange reaction in the presence of KNH2 is
very fast as compared to the gasliquid mass transfer rate (at the
temperature and catalyst concentration employed on industrial
scale). The rate of mass transfer, therefore, is the controlling
step in the overall process. To achieve the requisite high mass
transfer rates, on each tray of the exchange towers, a large

722

S. Balamurugan et al. / Chemical Engineering Science 63 (2008) 721 731

number of ejectors are provided. The use of the ejector trays


reduces the size of the column required for the operation and
thereby reduces the capital cost of the equipment.
To design such gasliquid contactors, it is necessary to establish relationships between geometry of the ejector, the operating
conditions, and the performance of the ejector. The important
design parameters for such contactors would be the entrainment rate, pressure drop across the entire ejector, hold-up of
the phases within the ejector, etc. Computational uid dynamics (CFD) modeling approach provides a better understanding
of the hydrodynamics of such systems. In this work, the effect of operating parameters like gas velocity, liquid level, and
ejector geometry (nozzle diameter and throat height) on liquid
entrainment rate and pressure drop are studied in detail with
the help of experiments and CFD modeling.

uid. The experiments were carried with an acrylic column of


0.06 m in diameter and 1 m in height. The airow rate was manually controlled with the help of a calibrated rotameter. The
air pressure just before the entry into the nozzle was measured
using a digital pressure gauge (AZ instrument make with accuracy of 0.3%). The entrainment rate of the liquid was measured
manually by collecting the liquid from the gasliquid separation tank in a known period. The ratio of the entrained water
to the amount of air is called as entrainment ratio (kg of water carried per kg of air). The velocity of air owing through
the nozzle was varied over a wide range. The corresponding
entrainment rate and pressure drop were measured for each airow rate. Table 1 shows the various geometry parameters that
were investigated. These experimental results were used for the
validation of CFD predictions.
The gas hold-up was calculated using the differential pressure measured between two points in the ejector. Pressures in
the ejector were measured using digital pressure gauge (AZ
instrument make with accuracy of 0.3%) at P1 (at a distance of
0.5 m from the nozzle tip) and P2 (at a distance of 0.8 m from

2. Experimental setup
A schematic diagram of the experimental setup is shown in
Fig. 1 with air as the motive uid and water as the entrained

Air outlet
Separator
Baffle
POUT
Pressure outlet boundary
condition (B.C.) (1 atm)
Straight tube/Column
Water
Pressure point P2

Two-phase
mixture

Water level, Pressure inlet (1 atm)


Pressure point P1

DC

Water
Diffuser
Liquid level
(LH)
HT

Throat
DT

Suction Tank
Converging section (inlet diameter = DEC)
Air mass flow rate B.C. at nozzle tip
Nozzle

DN
D0

PIN
Rotameter
Compressor
Air inlet

Valve
Fig. 1. Gasliquid ejector experimental apparatus for measuring liquid entrainment and boundary conditions used in CFD simulations.

S. Balamurugan et al. / Chemical Engineering Science 63 (2008) 721 731


Table 1
Dimensions of the ejector and parameters varied in the experiments
Parameter

Values

Nozzle inlet diameter D0 (m)


Nozzle diameter DN (m)
Throat diameter DT (m)
Throat height HT (m)
Column diameter Dc (m)
Column height (m)
Water level in the suction tank LH (m)
Pressure at D0 (N/m2 , gauge)
Airow rate (m3 /s)

0.0254
0.004, 0.006, 0.008, 0.01, 0.012
0.02, 0.0254, 0.04
0.05, 0.1
0.06
1
0.25, 0.3, 0.35, 0.4, 0.45, 0.5
1013114 000
0.00260.026

the nozzle tip). The local pressure values were acquired using
an AD card in a personal computer for every second for about
10 min. These values were time averaged to get local average
pressure and hence the pressure drop across the two points.
The differential pressure measurements were used to calculate the gas hold-up using Eq. (1)(4):
P = mixture gh.

(1)

The distance between the two pressure probes was 0.3 m.


Since the densities of air and water are known, the mixture
density can be dened by
mixture = L L + G G .

(2)

Since this is two-phase system, Eq. (3) can be used:


L + G = 1.

(3)

Using above equations the pressure drop can be related to the


gas hold-up as
P = {L (1 G ) + G G }gh.

(4)

3. CFD modeling strategy


Air emerging from the nozzle enters the ejector along with
water. The airwater mixture ows through the converging
section, throat, diffuser (diverging section), and straight tube.
The ow through the throat and the straight pipe region of
the ejector resembles that of two-phase ow in vertical pipe.
Hence, the CFD simulations were rst performed for the twophase ow in a vertical pipe. The mixture model was used in
the CFD simulations including the interfacial forces like drag
and lift forces. These forces depend on the droplet/bubble diameter of the dispersed phase. The experimental gas hold-up
measurements showed that the hold-up of liquid in the ejector
column was in the range of 0.10.25. Hence, all the simulations
were performed with the liquid phase dispersed in the form of
droplets in the continuous gas medium.
In the CFD simulations of gasliquid contactors like stirred
tanks and bubble columns, the droplet/bubble size of the
dispersed phase is a tted parameter or an assumed value.
For example, in bubble column simulations, Sanyal et al.
(1999) and Pan et al. (1999) have used a bubble diameter of

723

0.005 m. Peger and Becker (2001) have used a bubble diameter of 0.004 m. This value was considered to be constant
in all simulations. Similarly in stirred tank simulations, Deen
et al. (2002) have used 0.002 and 0.004 m bubble diameter and
Khopkar et al. (2005) have used 0.004 m constant bubble diameters. The droplet diameters in ejectors are generated by the
shear created near the nozzle region. Thus, the droplet diameter
will be different for different gas velocities.
The CFD simulations were carried out for two phase ow
in pipes. In the vertical pipe simulations, the droplet diameter
was varied so that the predicted pressure drop and the hold-up
match with the experimental data reported by Anderson and
Mantzouranis (1960) for a range of gas velocities. This enabled
us to establish a relationship between the droplet diameter and
the hydrodynamic conditions within the pipe and the gas velocity. This relationship (given later in the manuscript) has been
used in all further simulations of ejectors studied in this work.
The experimental data of Anderson and Mantzouranis (1960)
were used because the values of the gas hold-up reported in
their work ranged from 0.75 to 0.95, which are similar to those
observed in the present work. Thus, for all further ejector simulations, the droplet diameter is not a tted parameter making
the CFD simulations predictive in nature.
A 3D geometry of the pipe was created and meshed in
Gambit software with 0.2 million hexahedral structured cells.
More than the total number of cells, what is important is the
size of the grid in the region of high velocity gradients such
as the nozzle and the throat. In our previous work (Kandakure
et al., 2005, 2007), we have studied this aspect in detail. The
mesh size in the region close to the nozzle and the ejector wall
were small enough to capture the complex ow phenomenon.
During the initial stages of the simulations, course grids (0.08
million, 0.1 million) were used for simulation. In these simulations, the momentum and k. quantities did not converge to
the required convergence criteria. Simulations with 0.2 million
grid size gave good convergence. The entrainment rate and
liquid hold-up predicted using the CFD simulations with
0.2 million matched quite well with the experimental results.
All the CFD simulations were performed using Fluent 6.2 software. Since the ow is vertically upward in the pipe, the gravity
was taken in the negative Z direction and the operating pressure was taken as 1 atm. The standard k. model was used for
the modeling of turbulent behavior of the ow. FLUENT uses
its in-built slip velocity formulation given by Manninen et al.
(1996). The mass ow rates of air and water were given as the
inlet boundary condition. Since the pipe outlet is open to atmosphere, the outlet boundary was taken as 0 gauge pressure.
The no-slip boundary condition was enforced at the walls of
the pipe. The second order upwind discretization scheme was
used for the momentum, volume fraction, turbulent kinetic energy and turbulent energy dissipation rate and SIMPLE scheme
was used for the pressurevelocity coupling. For all variables
under-relaxation factor of 0.2 was used.
The CFD models used for the pipe simulations were used
to simulate the ejector. Fig. 1 also shows the boundary condition used in the ejector simulation. Air and water at room
temperature were considered as motive and entrained uids,

S. Balamurugan et al. / Chemical Engineering Science 63 (2008) 721 731

respectively. The air was assumed to obey an ideal gas law. The
mass ow rate of air, measured from experiments, was provided
at the nozzle tip as the boundary condition. The top liquid surface in the suction chamber was considered as the input with a
value of 0 (zero) gauge pressure. The ejector outlet is open to
atmosphere and, therefore, the pressure outlet boundary condition was used with 0 (zero) gauge pressure. For both the pipe
and ejector simulations, the solution was iterated until convergence was achieved, such that the residue for each equation fell
below 102 . In general, it was observed that the residue for
the momentum equations was below 103 ; that for the turbulent kinetic energy and turbulent energy dissipation rate were
well below 103 . The continuity equation residue was below
102 , while the residue for the liquid volume fraction was well
below 103 .

12000

Pressure drop, (N/m2)

724

8000

4000

0
0.000

0.002
0.004
Mass flow rate air, (kg/s)

0.006

0.002
0.004
Mass flow rate of air (kg/s)

0.006

4. Results and discussion


4.1. Vertical pipe simulations

dP = 1.778 10

VGT .

(5)

In this case, there are two major effects. As the gas velocity
increases, the shear rates in the throat increase. However, unlike other contactors, the liquid entrainment rate also increases
dramatically. This means that a larger quantity of liquid needs
to be dispersed in the gas. As a combined effect of these two
contrary effects, the liquid droplet diameter increases with the
gas velocity. In fact, the increase in liquid entrainment rate is
very large. There are no similar equations in the literature because his aspect has not been studied in the past literature. One
more point to be noted is that, CFD simulations were carried
out with mixture model in Fluent. In the mixture model, the interaction between the two phases is captured through the drag
forces acting on the droplets of the dispersed phase. The drag

Gas hold-up (-)

0.95

For the vertical pipe, at particular ow rates of air and water, the simulation was performed to predict pressure drop and
gas hold-up. As discussed earlier, the droplet diameter was varied until the predicted values of pressure drop and gas hold-up
matched with the experimentally reported values. Similar procedure was followed for other ow rates of air and water.
Figs. 2A and B show the comparison of the predicted values
of pressure drop and gas hold-up with the experimental measurements of Anderson and Mantzouranis (1960). The gures
show that when the airow rate is increased, the pressure drop
and gas hold-up in the pipe also increases. Fig. 2B shows that
the predicted values of the gas hold-up compare well with the
experimental values. However, the experimental pressure drops
are slightly on the higher side (Fig. 2A). This could be because
Anderson and Mantzouranis (1960) measured the pressure drop
including the pressure drop across the air rotameter. Hence, the
pressure drop indicated in the experiments is actually a sum of
the pipe pressure drop and the rotameter pressure drop (not the
region of interest).
The droplet diameter values tted from the above simulations
were related to the supercial gas velocity in the following
manner:

0.85

0.75

0.65
0.000

Fig. 2. (A) Comparison of CFD predicted pressure drop with experimental


results of Anderson and Mantzouranis (1960) for constant liquid ow rate of
0.0123 kg/s. : Experimental and : predicted. (B) Comparison of CFD
predicted gas hold-up with experimental results of Anderson and Mantzouranis
(1960) for constant liquid ow rate of 0.0123 kg/s. : Experimental and
: predicted.

force depends upon the relative velocity of the two phases as


well as the size of the dispersed phase droplets. The drag forces
increase with an increase in the size of the dispersed phase
droplet. Another interpretation of Eq. (5) is that as the gas velocity increases the momentum exchange between the gas and
the liquid increases. This can be captured through an increase
in the droplet size of the dispersed phase. Thus, Eq. (5) indicates the increased momentum transfer between gas and the
liquid phases (by the way of size of dispersed phase droplet)
as the gas velocity increases.
In order to test the efcacy of the above relationship, it was
decided to use the above equation to calculate drop size and use
this value for the CFD prediction for a different set of experimental data. For this purpose, the experimental data reported
by Gill et al. (1965) with vertical pipe diameter of 0.03175 m
and height of 1.82 m were used. In the simulation, water ow

S. Balamurugan et al. / Chemical Engineering Science 63 (2008) 721 731

725

0.5

0.4
Liquid hold-up (-)

Pressure drop (N/m2)

1200

800

0.3

0.2

400
0.1

0
0.00

0.02

0.04
0.06
Mass flow rate air (kg/s)

0.08

0.10

Gas hold-up (-)

0.96

120

140
160
Nozzle velocity (m/s)

180

200

Fig. 4. Comparison of liquid hold-up predicted by CFD with experimental results obtained with pressure measurements for DN =0.008 m, DT =0.0254 m,
HT = 0.1 m and LH = 0.5 m. : Experimental and : predicted.

0.92

0.88

0.84
0.00

0
100

0.02

0.04
0.06
Mass flowrate air (kg/s)

0.08

0.10

Fig. 3. (A) Comparison of CFD predicted pressure drop with experimental


results of Gill et al. (1965) for constant liquid ow rate of 0.025 kg/s. :
Experimental and : predicted. (B) Comparison of CFD predicted gas
hold-up with experimental results of Gill et al. (1965) for constant liquid
ow rate of 0.025 kg/s. : Experimental and : predicted.

rate of 0.025 kg/s and the airow rate range from 0.0125 to
0.0875 kg/s were used. Figs. 3A and B compare the pressure
drop and gas hold-up values predicted by CFD with experimental values reported by Gill et al. (1965). The gures show that
the CFD predictions match quite well with those of the experiments. Hence, the relation developed to estimate the droplet
diameter was able to predict the pressure drops and the hold-up
values with good accuracy.
4.2. Ejector simulations
Eq. (5) developed using the vertical pipe simulations was
used for the ejector simulation. The air jet emerging from the
nozzle enters the ejector through the converging section and
it carries the entrained liquid along with it. The dispersion

and formation of the liquid droplets occur mainly in the throat


section. Hence, the gas velocity at the throat was used in Eq.
(5) to predict the droplet diameter.
The gas velocity at the throat was estimated based on our
earlier work (Balamurugan et al., 2006). From the measured
value of gas mass ow rate and the inlet pressure (PIN ), the
pressure at the nozzle exit was calculated from the theory of
adiabatic compressible ow. Using this value of pressure, gas
density at the nozzle tip can be calculated. This enabled calculation of the gas velocity at the tip of the nozzle. The throat
velocity, VGT is calculated from the nozzle tip gas velocity, by
accounting for the area ratio and assuming that the gas density remains the same. Thus, for example, if the gas velocity at
the nozzle tip is 150 m/s and the area ratio is 0.1, the value of
VGT would be 15 m/s. The range of gas velocity in the throat is
about 10.20 m/s. The corresponding droplet diameter is about
1.2 mm. From the CFD simulations, the amount of entrained
liquid, hold-up proles and pressure proles were obtained.
The hold-up predictions are vital for validation of the hydrodynamic simulation of the ejectors. Fig. 4 shows the comparison
of gas hold-up values predicted from the CFD and the measured values. The hold-up values predicted from CFD are in
good agreement with that of the experimental. It is worthwhile
to reiterate that the droplet diameters have been tted to an experimental data set of Anderson and Mantzouranis (1960). This
tted droplet diameters have been related to the gas velocity.
This relationship between the droplet diameter and gas velocity
has been used for all the predictions in this work. This relationship successfully predicts not only the data for two-phase pipe
ow in the previous literature but also our own experimental
data and the experimental data on ejectors reported in the past
literature. Therefore, this relationship between droplet diameter and gas velocity can be considered to be extremely useful.
Similar predictions were performed for the ejectors and the effects of operating parameters like gas velocity and liquid level

726

S. Balamurugan et al. / Chemical Engineering Science 63 (2008) 721 731

Fig. 5B shows the variation of the gauge pressure along the


center line of the ejector with the axial location from the ejector
entry for different nozzle gas velocities. On top of the gure,
ejector has been shown. Since the ejector outlet was open to
atmosphere, the pressure at this location is 0 (zero) Pa gauge.
In the throat region, the two-phase ow causes a reduction in
pressure and at the exit of the throat, a minimum pressure zone
exists. This pressure is lower than the atmospheric pressure that
exists at the liquid level of the suction chamber and the difference between these two pressures becomes the driving force
for the liquid entrainment. In the diffuser section, some pressure recovery takes place due to the increase in cross-sectional
area. In the straight tube section, again the pressure reduces.
When the nozzle velocity was increased from 48 to 132 m/s, the
pressure at the throat exit decreased. This increases the driving
force for the liquid entrainment. Hence, the liquid entrainment
rate increases with an increase in the nozzle velocity.
Eq. (5) indicates that the relationship between the throat velocity and droplet size is linear. This means that if the throat
velocity is doubled, the droplet size will double. This will cause
enhanced momentum transfer from the gas phase to the liquid phase causing an increase in the liquid entrainment rate.
For example, when the nozzle velocity is increased from 80 to
160 m/s (correspondingly throat velocity also doubles), entrainment rate increases by about 1.6 times (Fig. 5A). This indicates
0.8 . This
that the entrainment rate is proportional to VN0.8 or VGT
means that the predictions are fairly sensitive to the values of
throat velocity.

0.7

QL (m3/hr)

0.6

0.5

0.4

0.3
75

125
Gas velocity (m/s)

175

5000

Pressure (Pa, gauge)

-5000

VG = 48 m/s
VG = 55 m/s

-10000

-15000
VG = 86 m/s
VG = 132 m/s

4.4. Effect of liquid level

-20000
0.1

0.3

0.5
0.7
Axial distance (m)

0.9

1.1

Fig. 5. (A) Effect of nozzle velocity on entrainment rate for DN = 0.008 m,


HT = 0.1 m, DT = 0.0254 m, LH = 0.5 m. : Experimental and :
predicted. (B) Effect of nozzle velocity on pressure prole from the centre line
of ejector with the axial locations from the nozzle inlet for DN = 0.008 m,
:
DT = 0.0254 m, LH = 0.50 m. : VG = 48 m/s;
VG = 55 m/s; - - - - - - - - - - - - -: VG = 86 m/s and
: VG = 132 m/s.

in the suction chamber and the geometry of ejector like area


ratio (area of nozzle to throat), and throat height on liquid entrainment are discussed in the following sections with the help
of both experimental and CFD predictions.
4.3. Effect of gas velocity
Fig. 5A shows the comparison between the predicted values
of liquid entrainment rate and the experimental measurements
for nozzle diameter of 0.008 m, throat diameter of 0.0254 m,
throat height of 0.1 m and liquid level of 0.5 m for different
nozzle velocities. The liquid entrainment rate increases with
the increase in nozzle velocity. The gure clearly indicates
that the predicted values of entrainment rate match well with
experimental values.

Fig. 6A shows the comparison between the liquid entrainment rate predicted from CFD simulations with the experimental measurements for the nozzle diameter of 0.008 m,
throat diameter of 0.0254 m, throat height of 0.1 m and mass
ow rate of 0.0054 kg/s for different liquid levels. The liquid
entrainment rate increases with the increase in liquid level.
Fig. 6A shows that the predicted values of entrainment rate
match well with the experimental values as points. Fig. 6B
shows the variation of pressure at throat exit for different liquid levels for the same mass ow rate of 0.0131 kg/s. When
the liquid level was increased from 0.3 to 0.5 m, the pressure
at the throat exit decreased. The momentum generated by the
gas jet and the liquid head provided by the liquid level are the
driving forces for the entrainment rate. Hence, when the liquid level is increased, the pressure at the throat exit decreases
increasing the liquid entrainment rate.
4.5. Effect of area ratio
The ratio of nozzle area (AN ) to throat area (AT ) (area ratio) is one of the important parameters for the ejector design.
Fig. 7A shows both the predicted and measured liquid entrainment rates at a constant gas mass ow rate of 0.0031 kg/s for
different area ratios ranging from 0.025 to 0.893 (nozzle diameters from 0.004 to 0.024 mm for a xed throat diameter of

S. Balamurugan et al. / Chemical Engineering Science 63 (2008) 721 731

0.60

QL (m3/hr)

0.40

0.20

0.00
0.3

0.4
Liquid level (m)

0.5

-14000

Pressure at throat (Pa)

-16000

-18000

-20000

-22000

727

gas jet. This causes a reduction in the annular area available for
water to ow (Fig. 7E). When the nozzle diameter approaches
the diameter of the throat, the area available for the liquid to
ow decreases. Water starts to climb on the outer walls of the
ejector due to reduction in the space available for entrained
liquid to ow into the ejector. Kandakure et al. (2005) have
reported similar optimum area ratio and shown a recirculation
pattern inside the ejector as the reason for reduction in the entrainment rate at higher nozzle diameters. It has been reported
that at high nozzle diameters (high value of AT /AN ), a substantial amount of recirculation occurs within the converging
section of the ejector. This is primarily due to the reduction in
available area for the ow of the entrained uid.
The optimum nozzle diameter (0.010 m) from the CFD predictions is in good agreement with the experimental results.
Several authors also have reported in the past that there is an
optimum value of the area ratio, however, no explanations were
provided by them. The optimum DN /DT ratio in this work corresponds to 0.393 and is in good agreement with the reported
values in the literature. Biswas and Mitra (1981) have reported
optimum DN /DT ratio to be in the range of 0.2230.258. Rylek
and Zahradnik (1984) have reported the optimum DN /DT ratio to be 0.33. Bando et al. (1990) have reported the optimum DN /DT ratio to be 0.50.l6, for a HT /DT ratio of 20.
Zahardnik et al. (1997) have reported that, as the nozzle diameter approaches the throat diameter, the entrainment rate decreases because the throat gets entirely lled with liquid for a
gasliquid system.
4.6. Effect of throat height

-24000

-26000
0.3

0.35

0.4
Liquid level (m)

0.45

0.5

Fig. 6. (A) Effect of liquid level on entrainment rate for DN = 0.008 m,


HT = 0.1 m, DT = 0.0254 m. : Experimental and : predicted. (B)
Effect of liquid level on pressure at throat exit for DN =0.008 m, HT =0.1 m,
DT = 0.0254 m.

0.0254 m). On increasing the area ratio from 0.025 to 0.155


(nozzle diameter from 0.004 to 0.010 m), the entrainment rate
increases. On further increasing the area ratio to 0.893 (nozzle diameter up to 0.024 m), the entrainment rate decreases.
Fig. 7B shows the variation of the gauge pressures along the
center line with axial locations from the ejector inlet for these
area ratios.
The CFD results (Figs. 7CE) show that the gas leaves from
the nozzle in the form of jet into the converging section of the
ejector, while the entrained liquid, ows through the annular
space near the wall. When the nozzle diameter increases from
0.004 to 0.010 m, the entrainment increases because the larger
diameter of the air jet increases the momentum transfer from
air to water. Fig. 7B shows that when the nozzle diameter is
increased from 0.004 to 0.010 m, the pressure at the throat exit
decreased and hence the driving force increases. However, a
further increase in the nozzle diameter increases the area of the

The effect of throat height of the ejector on the liquid entrainment rate was predicted with the help of the current model. It
was observed that the entrainment rate increases with decrease
in HT . Fig. 8 shows the effect of HT on the pressure prole
along the center line of the ejector. An ejector with no throat
was observed to have the maximum entrainment rate. This is
because the pressure value at the throat exit was smallest in this
case. This means the driving force for the suction that is the
pressure difference between the pressure at throat exit and atmosphere decreases with increase in throat height. Hence, the
entrainment rate decreases with increase in throat height. Similar results were reported by Kandakure et al. (2005) from the
CFD simulations of liquid jet ejectors. Henzler (1983), from
his review work on single phase horizontal ejectors, has given
the optimum ejector dimensions. He has observed that the optimum HT /DT ratio is 7.5 for (AT /AN ) of less than 2.5. Similar observations were made by Rylek and Zahradnik (1984).
The (HT /DT ) ratio used in Rylek and Zahradnik (1984) experiments ranged from 1.25 to 3.9. Bando et al. (1990) reported
that the optimum range of HT /DT ratio is 2030. Similar to the
present work results, Dirix and van der Wiele (1990) have reported that when throat height was increased, a decrease in the
efciency of the ejector was observed due to less entrainment
rate of gas. Havelka et al. (1997) have reported that the entrainment rate of gas increased with an increase in throat height.
However, they have also observed that the further increase in

728

S. Balamurugan et al. / Chemical Engineering Science 63 (2008) 721 731

Entrainment rate (m3/hr)

0.40

0.30

0.20

0.10

0.00
0.0

0.2

0.4

0.6

0.8

1.0

AN/AT

4000

Pressure, (Pa, gauge)

0
-4000
AR = 0.893 (DN = 0.024 m)

-8000

AR = 0.223 (DN = 0.012 m)

-12000

AR = 0.025 (DN = 0.004 m)

-16000

AR = 0.155 (DN = 0.010 m)

-20000
0.1

0.3

0.5
0.7
Axial distance (m)

0.9

1.1

Fig. 7. (A) Effect of area ratio (AN /AT ) on entrainment rate and throat pressure drop for LH = 0.5 m, HT = 0.1 m, DT = 0.0254 m and gas
mass ow rate = 0.0054 kg/s. : Experimental and : predicted. (B) Effect of area ratio (AN /AT ) on pressure at throat exit for LH = 0.5 m,
:
HT = 0.1 m, DT = 0.0254 m and gas mass ow rate = 0.0054 kg/s. : AR = 0.025 (DN = 0.004 m);
AR = 0.155 (DN = 0.010 m); - - - - - - - - - - - - -: AR = 0.223 (DN = 0.012 m) and : AR = 0.893 (DN = 0.024 m). (C) Predicted liquid volume fraction prole for LH = 0.5 m, HT = 0.1 m, DT = 0.0254 m and gas mass ow rate = 0.0054 kg/s for DN = 0.004 m.
(D) Predicted liquid volume fraction prole for LH = 0.5 m, HT = 0.1 m,
0
0.5
1.0
DT = 0.0254 m and gas mass ow rate = 0.0054 kg/s for DN = 0.010 m.
(E)
0
0.5
1.0
Predicted liquid volume fraction prole for LH = 0.5 m, HT = 0.1 m, DT = 0.0254 m and gas mass ow rate = 0.0054 kg/s for DN = 0.024 m.

0.5

1.0

the entrainment became almost negligible for HT /DT greater


than 6.
5. Predictions for other geometries
The CFD model developed for the ejector was validated with
the experimental results obtained in this work. However, in

order to nd the predictive ability of the model to other geometries, it was decided to predict the performance of ejectors
reported in the past literature. For this purpose, the ejector geometry reported by Davies et al. (1967) was used for the CFD
simulations. The liquid entrainment rates reported by Davies
et al. (1967) were compared with the CFD predictions. Davies
et al. (1967) have preformed experiments with a gas jet ejector

S. Balamurugan et al. / Chemical Engineering Science 63 (2008) 721 731

729

Pressure outlet

HT = 0.2

HT = 0.1

Pressure (Pa, gauge)

HT = 0

0
0.1

0.3

0.5

0.7

0.9

1.1

DT = 1.56

-20000

-40000

DC = 0.0381

-60000
Axial distance (m)
Fig. 8. Effect of throat height on pressure prole from the center line of
ejector with the axial locations from the nozzle inlet for DN = 0.008 m,
:
DT = 0.0254 m, LH = 0.50 m. - - - - - - - - -: HT = 0;
HT = 0.1 and : HT = 0.2.

0.2667 m
DT = 0.0127
DN = 0.00095, 0.00133
Mass flow inlet

Water inlet,
Diameter = 0.01905 m

Recycled water inlet


Suction Chamber
Air

0.14
Liquid Entrinament rate (kg/s)

with two different nozzle diameters (0.0009 and 0.00133 m)


and throat diameter of 0.0127 m. The ejector dimensions used
by Davies et al. (1967) have been given in Fig. 9A.
Fig. 9B shows the comparison of liquid entrainment rate predicted from the CFD simulations with the reported values. The
gure shows that the predicted values of liquid entrainment rate
match with the reported values very well. Hence, the present
CFD simulation can be considered to have good predictive
capability.
The CFD, therefore, enables the prediction of local pressures
in the ejectors, which are otherwise very difcult to measure.
Fig. 10 shows the plot of liquid entrainment rate plotted against
the absolute value of pressures at the throat exit.
The results show that the ratio of the nozzle diameter to the
throat diameter plays a crucial role in determining the hydrodynamics and the performance of the ejectors. Fig. 11A shows
the effect of nozzle diameter on the entrainment ratio (L/G, ratio of amount of water entrained to the amount of air supplied)
as a function of the air mass ow rate. This gure shows that
for a particular nozzle diameter, the L/G goes through a maximum. This is due to the leveling of entrainment rate observed at
higher gas velocities (Fig. 5A). The maximum L/G ratio is observed for the nozzle diameter of 0.010 m as explained earlier.
This means that for a given production rate in the plant (given
mass ow rate) the extent of gasliquid contacting (L/G ratio) can be increased simply by changing the nozzle size. This
also has an added advantage. Fig. 11B shows the effect of nozzle diameter on the pressure drop. The total pressure drop in
Fig. 11B refers to the pressure drop across entire ejector (between the inlet pipe to the nozzle and the gas outlet from the
ejector). Since the gas outlet from the ejector is to atmosphere,

0.159 m

0.12
AR = 0.01
0.10

AR = 0.0056

0.08
0.06
0.04
0.02
0.00
0

50

100
150
Gas velocity (m/s)

200

250

Fig. 9. (A) Experimental setup used by Davies et al. (1967) to measure


liquid entrainment rate. (B) Comparison of CFD predicted pressure drop
with experimental results of Davies et al. (1967) for DT = 0.0127 m. :
AR = 0.0056 (DN = 0.0009 m); : AR = 0.01 (DN = 0.0013 m) and :
predicted.

730

S. Balamurugan et al. / Chemical Engineering Science 63 (2008) 721 731

Liquid Entrainment rate (m3/hr)

0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
0

5000
10000
15000
Pressure at Throat Exit (Pa)

20000

the total pressure drop across the ejector is the same as the
pressure before the nozzle inlet (PIN ). This was measured experimentally (the measurement location has been indicated in
Fig. 1). It can be seen from the gure that, for the same mass
ow rate, when the nozzle diameter was increased from 6 to
12 mm, there was a substantial reduction in the pressure drop.
For example, for a mass ow rate of 0.0075 kg/s, an increase
in the nozzle diameter from 6 to 12 mm leads to a reduction in
the pressure drop from 1.5 105 to about 1.05 105 N/m2 .
This reduction is close to 30%. Combination of Figs. 11A and
B shows that for a given mass ow rate of air, the extent of
contacting can be increased (higher L/G ratio) and simultaneously operating cost can be reduced (lower pressure drop) by
using an optimized value of nozzle diameter.
6. Conclusions

Fig. 10. Liquid entrainment rate correlation with pressure at the throat exit.

40
10 mm

L/G

30

20
4mm

10

6 mm

8 mm
12 mm

0
0.000

0.006
0.012
Gas mass flow rate (kg/s)

0.018

2.0E+05

Total pressure drop (Pa)

6 mm
4 mm

1.5E+05

8 mm

Notation

10 mm

12 mm

1.0E+05
0.000

0.006
0.012
Gas mass flow rate (kg/s)

In the present work, hydrodynamic characteristics of ejectors


using air as the motive uid and water as the entrained uid
have been investigated. Experiments have been performed over
a wide range of ejector congurations (DN = 0.004.0.012 m,
DT = 0.02.0.04 m, HT = 0.05 and 0.1 m, nozzle velocity =
27.210 m/s and liquid level = 0.25.0.50 m. A CFD model was
developed and the performance of the ejector was predicted.
From a particular data of two-phase up-ow system, a relation
was developed to predict the droplet diameter. This relation
was used for all further simulations and it was observed that
the CFD predictions agree well with the experimental values.
It was observed that the liquid entrainment rate increases with
an increase in the liquid level and the gas velocity. The entrainment rate was found to be the highest corresponding to area
ratio of about 0.155. This is because, at lower area ratio, the
momentum transferred from air jet is low and at higher area ratio, area available for water to ow decreases. In addition, the
entrainment rate reported by Davies et al. (1967) has also been
predicted with the model developed. It was also observed that
the absolute value of pressure at throat exit and the liquid entrainment rate correlates well. Hence, by optimizing the nozzle
diameter, the extent of contacting can be increased (increase
in L/G by about 50%) along with reduction in operating cost
(reduction in pressure drop of about 30%) can be achieved.

0.018

Fig. 11. (A) Effect of nozzle diameter on L/G ratio for constant DT =0.0254 m,
HT = 0.1 m, LH = 0.5 m. : DN = 0.004 m; :
: DN = 0.010 m
DN = 0.006 m; : DN =0.008 m;
and - - - - - - - - -: DN = 0.012 m. (B) Effect of nozzle diameter on total pressure drop across the ejector for DT = 0.0254 m, HT = 0.01 m and
LH = 0.5 m. : DN = 0.004 m; : DN = 0.006 m; : DN = 0.008 m; :
DN = 0.010 m; and : DN = 0.012 m.

AR
dP
DC
DN
DT
g
h
HT
LH
P
PIN
POUT

ratio of nozzle area to throat area


droplet diameter, m
diameter of column, m
diameter of nozzle, m
Diameter of throat, m
acceleration due to gravity, m/s2
distance between two pressure points, m
height of the throat, m
liquid level in the suction tank, m
pressure drop, N/m2
pressure at nozzle inlet, N/m2
pressure at ejector outlet, N/m2

S. Balamurugan et al. / Chemical Engineering Science 63 (2008) 721 731

VG
VGT

gas velocity at the nozzle, m/s


gas velocity at the throat, m/s

Greek letters
G
L
G
L
mixture

volume fraction of gas


volume fraction of liquid
gas density, kg/m3
liquid density, kg/m3
liquid density, kg/m3

Acknowledgment
The authors would like to acknowledge the nancial support
in the form of research grant & fellowship from the Department
of Atomic Energy (DAE), India (Project no. 47.01).
References
Anderson, G.H., Mantzouranis, B.G., 1960. Two phase (gasliquid) ow
phenomenaI, pressure drop and hold-up prole for two-phase ow in
vertical tubes. Chemical Engineering Science 12, 109126.
Balamurugan, S., Gaikar, V.G., Patwardhan, A.W., 2006. Hydrodynamic
characteristics of gasliquid ejectors. Chemical Engineering Research and
Design 84, 11661179.
Bando, Y., Kuraishi, M., Nishimura, M., Takeshita, I., 1990. The
characteristics of a bubble column with a gas-suction type, simultaneous
gasliquid injection-nozzle. International Chemical Engineering 30,
729737.
Ben Brahim, A., Prevost, M., Bugarel, R., 1984. Momentum transfer in
a vertical down ow liquid jet ejector: case of self gas aspiration and
emulsion ow. International multiphase ow 10 (1), 7994.
Biswas, M.N., Mitra, A.K., 1981. Momentum transfer in horizontal multijet liquidgas ejector. Canadian Journal of Chemical Engineering 59,
634637.
Cramers, P.H.M.R., Beenackers, A.A.C.M., 2001. Inuence of the ejector
conguration, scale and the gas density on the mass transfer characteristics
of gasliquid ejectors. Chemical Engineering Journal 82, 131141.
Davies, G., Mitra, A.K., Roy, A.N., 1967. Momentum transfer studies
in ejectors. Industrial and Engineering Chemistry Process Design and
Development 6 (3), 293299.

731

Deen, N.G., Solberg, T., Hjertager, B.H., 2002. Flow generated by an aerated
Rushton impeller: two-phase PIV experiments and numerical simulations.
Canadian Journal of Chemical Engineering 80 (4), 638652.
Dirix, C.A.M.C., van der Wiele, K., 1990. Mass transfer in jet loop reactors.
Chemical Engineering Science 45, 23332340.
Gill, L.E., Hewitt, G.F., Lacey, P.M.C., 1965. Data on the upwards annular
ow of airwater mixtures. Chemical Engineering Science 20, 7188.
Havelka, P., Linek, V., Sinkule, J., Zahradnik, J., Fialova, M., 1997. Effect
of the ejector conguration on the gas suction rate and gas hold-up in
ejector loop reactors. Chemical Engineering Science 52, 17011713.
Havelka, P., Linek, V., Sinkule, J., Zahradnik, J., Fialova, M., 2000.
Hydrodynamics and mass transfer characteristics of ejector loop reactors.
Chemical Engineering Science 55, 535549.
Henzler, H., 1983. Design of ejectors for single phase material systems.
German Chemical Engineering 6, 292300.
Kandakure, M.T., Gaikar, V.G., Patwardhan, A.W., 2005. Hydrodynamic
aspects of ejectors. Chemical Engineering Science 60, 63916402.
Kandakure, M.T., Patkar, V.C., Patwardhan, A.W., 2007. Characteristics
of turbulent conned jets. Chemical Engineering Processing, in press,
corrected proof available online 6 April 2007.
Khopkar, A.R., Rammohan, A.R., Ranade, V.V., Dudukovic, M.P., 2005.
Gasliquid generated by a Rushton turbine in stirred vessel: CARPT/CT
measurements and CFD simulations. Chemical Engineering Science 60,
22152229.
Manninen, M., Taivassalo, V., Kallio, S., 1996. On the mixture model for
multiphase ow. VTT Publications 288, Technical Research Centre of
Finland.
Mukherjee, D., Biswas, M.N., Mitra, A.K., 1988. Hydrodynamics of
liquidliquid dispersion in ejectors and vertical two phase ow. The
Canadian Journal of Chemical Engineering 66, 896907.
Pan, V., Dudukovic, M.P., Chang, M., 1999. Dynamic simulation of bubbly
ow in bubble columns. Chemical Engineering Science 54, 24812489.
Peger, D., Becker, S., 2001. Modelling and simulation of the dynamic
ow behaviour in a bubble column. Chemical Engineering Science 56,
17371747.
Rylek, M., Zahradnik, J., 1984. Design of Venturi-tube gas distributors
for bubble type reactors. Collection of Czechoslovak Chemical
Communications 49, 19391948.
Sanyal, J., Vasquez, S., Roy, S., Dudukovic, M.P., 1999. Numerical simulation
of gasliquid dynamics in cylindrical bubble column reactors. Chemical
Engineering Science 54 (21), 50715083.
Zahradnik, J., Fialova, M., Linek, V., Sinkule, J., Renickoca, J., Kastanek,
F., 1997. Dispersion efciency of ejector-type gas distributors in different
operating modes. Chemical Engineering Science 52 (24), 44994510.

You might also like