You are on page 1of 23

ANALYSIS OF GAS CARRY-UNDER IN GAS-LIQUID CYLINDRICAL CYCLONES

by

S. K. Marti, F. M. Erdal, O. Shoham, S. A. Shirazi


The University of Tulsa
and
G. E. Kouba
Chevron Petroleum Technology Company

ABSTRACT

Existing theories for cyclone separators, developed for liquid-liquid, liquid-solid and gas-solid
flows, can handle low concentrations of the dispersed phase. These models are not appropriate
for gas-liquid cylindrical cyclone (GLCC) separators over a full range of gas fractions, because
the flow dynamics and phase distribution are not limited to dispersed flow. Indeed, in the gas-
liquid cyclone several flow patterns may occur, including churn, mist, annular, dispersed bubbles
and separated flows. Furthermore, several of these flow patterns may occur in the cyclone under
the same inlet flow conditions.

This paper presents a first attempt to develop a mechanistic model to predict gas carry-under in
GLCC separators. The model predicts the gas-liquid interface near the GLCC inlet as a function
of the radial distribution of the tangential velocity. The interface defines the starting location for
the bubble trajectory analysis, which enables determination of gas carry-under and separation
efficiency for the GLCC. The decay of the tangential velocity in the axial flow direction is
incorporated. Also presented are preliminary CFD simulation results obtained from a commercial
package (CFX). The proposed study is a part of a comprehensive model for a proper design of
industrial GLCC separators.
INTRODUCTION

Cyclone separators have been in existence for at least 150 years, yet the use of cyclones in the oil
field is a fairly recent phenomenon. The high performance and relatively small size of cyclone
separators makes them an extremely attractive alternative technology for offshore applications.
The short fluid residence times of cyclones translate into much smaller equipment than
conventional gravity based separation facilities. This can be critical on offshore platforms, where
weight and space requirements have a significant multiplier effect on the overall cost of the
separation facilities. Liquid-liquid and liquid-solid hydrocyclone technology have grown rapidly
over the last decade, but full range compact gas-liquid cyclone technology is only just beginning
to emerge.

One of the promising cyclone technologies is the gas-liquid cylindrical cyclone (GLCC) separator.
The basic GLCC configuration is comprised of a cylindrical pipe section with a multiphase
tangential inlet, an upper outlet for gas, and a lower outlet for liquid. The simplicity of the GLCC
allows it to be easily and inexpensively manufactured, installed, and operated. A multiphase
metering loop as shown in Fig. 1 is one of the simplest arrangements for the GLCC. The self-
regulating nature of the loop allows the GLCC to operate over a wide range of gas and liquid
flow rates without additional liquid level controls. The GLCC is especially well suited for
applications requiring only partial gas-liquid separation. Examples are: control of gas-liquid ratio
upstream of multiphase pumps, multiphase meters and de-sanders, and pre-separation upstream of
slug catchers or primary separators. One of the main obstacles to the widespread deployment of
GLCCs has been the lack of good performance prediction and design tools.

Lg2

D3
Lg1

Lg3
P1

Lin Ll
w
P2
Ll1
Ll 3

D2

D1 Ll2a Ll2m Ll2b

Fig 1: GLCC Loop Nomenclature for Mechanistic Model

2
The state-of-the-art in predicting cyclone performance has in the past revolved solely around
computational fluid dynamic modeling of the flow field and the trajectories of discreet particles of
the dispersed phase. While well suited for local modeling of single phase and dispersed two-phase
flows, present CFD models are unable to handle some of the complex flow regimes observed in
the GLCC, in particular, slug and churn flows. Furthermore, CFD models of large system, e.g.,
the multiphase metering loop shown in Fig. 1, are too unwieldy to be practical for design
purposes.

Mechanistic modeling offers a practical approach to GLCC design and performance prediction
that is complementary to CFD work. Mechanistic models make simplifying assumptions while
still capturing the fundamental physics of the problem, consequently, they are not as detailed,
rigorous, or accurate as the CFD models. Advantages of mechanistic models include speed,
ability to model entire system, run on PC and, therefore, are more accessible to engineers as a
design tool than CFD models. Our approach is to develop fundamental mechanistic models and
verify and refine them with experimental data and CFD predictions.

Most of the studies published on cyclone separators have been focused on liquid-liquid, liquid-
solid and gas-solid flows. Very few studies on gas-liquid flows in the GLCC have been reported.
Recently, Kouba et al. (1995) and Arpandi et al. (1995) developed and discussed mechanistic
models that predict liquid carry-over with the gas stream. A comprehensive review of the
literature on gas-liquid flow in GLCCs and related topics is also given by them. Kurokawa and
Ohtaki (1995) conducted an experimental study on the characteristics of gas-liquid swirling flow
and the gas separation efficiency of a spiral type cyclone separator. A honeycomb type swirl
breaker is utilized to improve the gas separation efficiency.

Two studies for fluid system other than gas-liquid are relevant to the determination of separation
efficiency in a GLCC. Wolbert et al. (1995) utilized droplet trajectory analysis for the estimation
of separation efficiency for liquid-liquid (oil-water) flow in hydrocyclones. Trajectories were
characterized through a differential equation combining models for the three velocity
distributions, namely, axial radial and tangential, and the settling velocity defined by Stokes law.
A characteristic droplet diameter, d100, was defined which corresponds to the smallest droplet
diameter that will be separated with a 100% efficiency. Kutepov et al. (1978) conducted an
experimental and theoretical investigation on the separation efficiency of hydrocyclones for solid-
liquid separation. A method for calculating the amounts of solid phase carried with the liquid
phase in both the clear and thickened products of separation was developed based on stochastic
theory of separation processes.

The literature reveals a lack of both experimental studies and predictive methods for gas carry-
under phenomenon in GLCCs. The aim of this work is to lay the foundation for characterizing
gas carry-under and determining separation efficiency.

Three mechanisms have been identified as possible ways that gas is carried under with liquid: 1)
the trajectory of individual bubbles is too shallow for small bubbles to escape to the gas core, 2)
gas core instability results in the gas core filament whipping helically and occasionally breaking

3
off, and 3) a bubble swarm instability occurs with a sudden increase of liquid rate producing a
cloud of bubbles that are unable to migrate to the gas core. This study addresses the first
mechanism, i.e., the trajectory of small individual bubbles.

MECHANISTIC MODELING

The trajectory of an individual bubble is determined from a force balance on the bubble. This first
requires an approximation of the velocity field in the GLCC. The axial release point for the
bubble is determined from the location of the gas vortex. The d100 bubble is the smallest bubble
whose trajectory will just allow all bubbles of this size to be captured and flow upwards with the
gas core. The bubble capture efficiency is the percentage of bubbles of a certain size that are
captured in the gas core. In order to determine an overall separation efficiency with respect to
gas carry-under, the bubble size distribution and total amount of entrained gas would have to be
known. Although beyond the scope of the present paper, the overall separation efficiency is the
desired outcome of this work.

The model for gas carry-under and determination of separation efficiency is developed through
bubble trajectory analysis, by extending the studies of Kouba et al. (1994) and Arpandi et al.
(1995), and using a similar approach presented by Wolbert et al. (1995) for liquid-liquid flow in
hydrocyclones.

A schematics of the physical system is given in Fig. 1. The GLCC is shown in a multiphase flow
metering loop configuration. The gas and liquid mixture is introduced into the GLCC through an
inclined tangential inlet. The inclined section promotes stratification and pre-separation of the
phases, which then enter tangentially into the GLCC through a slot. The flow from the inlet slot
generates a swirl near the inlet of the GLCC. Due to the centrifugal forces, the liquid phase is
pushed to the wall while the gas phase moves to the center. The gas phase exits from the top into
the gas leg, and the liquid phase exits from the bottom through the liquid exit. A thin gas core
filament is created below the vortex by the smaller bubbles separated below the vortex region. In
this configuration the gas and the liquid phases are metered utilizing conventional flow meters
before being recombined at the exit of the loop. The figure includes all the dimensions of the
GLCC and the different sections of the flow loop.

Preliminary Calculations
Prior to bubble trajectory analysis, it is essential to determine the distribution of the gas and the
liquid in the GLCC. This includes the equilibrium liquid level and the gas-liquid interface shape
and location. These models were presented in detail by Arpandi et al. (1995) and are summarized
briefly below for completeness.

Equilibrium Liquid Level: The equilibrium liquid level simply corresponds to the pressure drop in
the GLCC as measured by an external sight gauge (glass level indicator). Because the frictional
losses are low in the GLCC, the equilibrium liquid level indicates the amount of liquid in the
GLCC.

4
The equilibrium liquid level is determined from a simple pressure balance between the inlet (P1)
and the outlet (P2) of the flow loop, for the gas leg and the liquid leg. This simple model neglects
any hydrodynamic interaction between the phases. Equating the pressure drops along the liquid
and gas legs one can solve explicitly for the equilibrium liquid level, as follows:

l g + l gLl3 g g ( Lin + Lg1 Lg3 )


Ll1 = (1)
l vl 2 f
g ( l g ) 1 l1
2 D1

where l and g are the frictional pressure losses in the liquid and gas sections, given as:

l n f i Li vi 2 m
l = + Kv 2
(2)
2 i =2 Di j =1
i i
l

g n f i Li vi 2 m
g = + Kv 2
(3)
2 i =1 Di j =1
i i
g

The terms inside the parentheses of Eqs. (2) and (3) represent the frictional losses in the different
pipe segments of the flow loop and the losses in the piping fittings, respectively.

Gas-Liquid Interface: The main assumption for this model is that the interface occurs due to a
forced vortex caused by the tangential inlet flow into the GLCC. The occurrence of a forced
vortex in a GLCC was substantiated by the experimental measurements of Millington and Thew
(1987). The tangential velocity distribution in the vortex is given by:

n
r
vt ( r ) = vt is (4)
Rs

where n = 1 for forced vortex, i.e., solid body rotation, n = -1 for free vortex, and -1<n<1 for a
combination vortex. In this study it is assumed that the tangential flow from the inlet generates a
forced vortex in the GLCC (n = 1). The tangential velocity from the inlet slot, vt is , is determined
from the liquid velocity at the inclined inlet section vlin , corrected for the slot area reduction and
for the inlet section inclination angle, ( = 270),as follows:

Ain
vt is = vlin cos( ) . (5)
Ais
where the liquid velocity at the inlet section, vlin , is calculated from the Taitel and Dukler (1976)
model.

5
The centrifugal force gives rise to a radial pressure difference across the vortex, namely, between
the gas-liquid interface and the GLCC wall, as follows:

r
l [
v t ( r )]2
P(r ) = dr . (6)
Rs r

The radial pressure difference is balanced by the hydrostatic pressure difference between the
liquid and the gas phases in the vortex. Thus, the axial location of the interface at any radial
position, r, can be obtained from the equation below:

P( r )
z( r ) =
( )
. (7)
g l g

The interface location z(r) can be used to calculate the total liquid volume displaced by the gas
vortex and gas core filament, yielding:

Rs

V g = 2rz (r )dr + Rc2 Llw z t ( ) (8)


Rc

where Rc is the gas core filament radius and zt = z(Rc) is the length of the vortex. The highest
point of the liquid vortex, where the interface touches the GLCC wall (vortex crown), can be
calculated assuming the total gas volume is submerged into the volume of liquid resulting from
the equilibrium liquid level calculation, as follows:

Vg
Llw = Ll1 + (9)
As

Bubble Trajectory
The starting axial location for the bubble trajectory is the bottom of the vortex. In the vortex
region the large bubbles are captured easily. At the bottom of the vortex the remaining small
bubbles are assumed to be homogeneously distributed. Below the vortex, the bubbles move
radially towards the GLCC centerline due to centripetal forces, and axially downward due to drag
forces from the axial flow of the liquid phase. If a bubble travels sufficiently radially inward, it
will merge with the gas core and will be carried upward by the gas stream. If, however, the
radial distance traveled by the bubble is insufficient, it will be carried under by the liquid stream
into the liquid leg exit.

Fig. 2 shows schematically the bubble trajectory model. The bubble is shown at time t and at time
t + t. The bubble moves radially at a velocity vr(r) and axially at the surrounding fluid

6
velocity, vz. The buoyancy force and slippage acting on the bubble in the z direction are neglected.
As a result the bubble will move farther in the radial direction, yielding an earlier gas carry-under
prediction, which is conservative from the design point of view. As a first approximation, the
velocity profile in the z direction is assumed to be uniform, namely, vz(r) = vz. The total
velocity of the bubble is vb(r), as shown in the figure. During the time interval, t, the bubble
moves dr = vr(r)t and dz = vz t in the radial and axial directions, respectively. Equating the
time period for the radial and axial movements of the bubble, and solving for the axial distance
yields the equation governing the bubble trajectory:

dr
dz = v z (10)
v r (r )

Interface

r
v(r)

z vz
vb
dz
dr
Trajectory

Fig 2: Schematics of Bubble Trajectory

The velocity distribution in the radial direction can be determined from a force balance on the
bubble. The forces acting on the bubble in the radial direction are the centripetal (centrifugal and
buoyancy), turbulent and drag forces.

Assuming local equilibrium at any radial location along the trajectory path traveled by the bubble,
a force balance on the bubble yields:

Frd = Frc Frt (11)

7
The centripetal force, Frc, is equal to the sum of the centrifugal force and buoyancy force acting
on the bubble, namely:

vt ( r ) 4
2

(
Frc = l g ) r 3
Rb 3 . (12)

The total drag force on the bubble, Fd, is determined from:

C v (r ) Rb
2 2

Fd = d l b (13)
2

and the radial component of the drag force is:

C v (r )vr (r )Rb
2
v
Frd = Fd cos = Fd r = d l b (14)
vb 2

The total turbulent force acting on the bubble is estimated in the same method proposed by Levich
(1962), as follows:

l vb' ( r ) Rb
2
2

Ft = (15)
2

where vb' ( r ) is the fluctuating component of the bubble velocity calculated by:

vb' ( r ) = vb2 f l
2
(16)

where fl is the friction factor. Substituting Eq. (16) into (15) and solving for the radial component
of the turbulent force yields:

fl vr (r )vb (r )Rb 2
Frt = (17)
2

Equating the radial forces from Eqs. (12), (14) and (17), as implied in Eq. (11), the radial velocity
distribution can be solved as:

4 l g v t (r )
2
db 1
vr (r ) = . (18)
3 l r (Cd f l ) vb (r )

8
The value of the drag coefficient, Cd, is calculated with a modified correlation developed by
Turton and Levenspiel (1986) and presented by Karamanev and Nikolov (1992) as:

Cd ( r ) =
[ . Re ( r )
24 1 + 0173
0.657
]+ 0.413
Re (r ) ( 1.09 )
1 + 16,300 Re (r )
for Re 400 (19a)

It was found out that the original correlation given in Eq. (19a) is valid in the range Re<400. A
new correlation is developed for the higher Reynolds number range, based on the data of
Haberman and Morton (1953), reported by Perry (1963), as follows:

Cd (r ) = 10 7 Re (r ) 2 + 10 3 Re (r ) + 01222
.
for 400 Re 5000 (19b)

Cd (r ) = 2.4 for Re 5000 (19c)

The Reynolds number is:

l vb ( r )d b
Re ( r ) = . (20)
l

The radial velocity distribution given by Eq. (18) can be substituted into Eq. (10). Integration of
Eq. (10) yields the bubble trajectory z(r), as follows:

r v
z( r ) = z dr . (21)
Rs v r ( r )

Stepping through r, from r = Rs to r = Rc maps out the entire trajectory of the bubble.

Tangential Velocity Decay


The maximum tangential velocity occurs at the inlet slot of the GLCC. As the flow swirls around
and moves downwards, the tangential velocity reduces due to viscous dissipation. This
phenomenon must be predicted and accounted for as it reduces the separation efficiency of the
GLCC.

The analysis is carried out on the swirling flow in a control volume bounded by the wall, as
shown in Fig. 3. A force balance on the control volume in the direction yields:

F = w dA cos = dVl l a (22)

9
where is the angle of the tangential velocity vector with respect to the horizontal. The wall
shear stress w is given by:

l v t2
w = f l (23)
8

dz

rd

dr

Fig 3: Control Volume for Tangential Velocity Decay Analysis

Substituting for the shear stress into Eq. (22), dividing through by l, and assuming that the
dVl dVl
characteristic length , is proportional to the radial increment dr, namely, = kdr , results:
dA dA

v t2
fl cos = kdra (24)
8

The acceleration is:

dv
a = v (25)
rd

Substituting Eq. (25) into Eq. (24) yields:

dv v t2
v = fl cos (26)
rd 8kdr

10
v
Dividing Eq. (26) by v2 , and recalling that = cos results in:
vt

dv fl
= rd (27)
v 8kdr cos

dz dz L
Substituting rd = and approximating = Eq. (27) becomes:
tan 2 kdr D

dv fl L
= (28)
v 4 sin D

As an example (used later in the CFD analysis), for v t is = 10 ft/s and vz = 0.5 ft/s, the angle of the
tangential velocity vector is = 30. Also, the smooth pipe friction factor for fully developed
dv L
turbulent flow is fl 0.06. For this case the tangential velocity decay is = 5.6% . This
v D
appears to be in a reasonable agreement with the experimental data of Hargreaves and Silvester
(1990) reported by Wolbert et al. (1995), and as shown later is in favorable agreement with the
present study CFD calculations.

Separation Efficiency
A schematics of the method for the determination of the separation efficiency is shown in Fig. 4.
As shown, the d100 bubble is the minimum bubble diameter which if released at the pipe wall (r
= Rs) below the vortex will move sufficiently radially to be captured by the gas core just before
the liquid outlet at the bottom of the GLCC. All bubbles of equal or larger diameter than the d100
bubble will be captured with an efficiency of 100%. However, smaller diameter bubbles, db< d100,
can be captured by the gas core only if they are released at a different radial location (Rc<r<Rs)
and not at the pipe wall. For this case, however, the probability of capturing the smaller bubble is
less than that of the d100 bubble, resulting in lower efficiency for this bubble size. For the example
shown in Fig. 4, the smaller bubble is released at the radial location re. This bubble will be
captured at any radial release location r<re, resulting in an effective area for capture of re2 . The
separation efficiency for this bubble is based on the ratio of the effective area for capture to the
total cross sectional area, namely:

2
re2 re
E (%) = = (29)
Rs2 Rs

11
Rs
re

Gas
Inlet
Flow

zt

Llw
d100 Lt

Liquid

Fig 4: Schematics of Separation Efficiency Analysis

RESULTS AND DISCUSSION

Mechanistic Model
The results obtained from the proposed model are plotted in Figs. 5-10. The results are predicted
for a GLCC metering loop, as shown schematically in Fig. 1, which was utilized in this study.
The GLCC is 3 in. in diameter, 7.6 ft high (3.4 ft and 4.2 ft above and below the inlet,
respectively). The inlet is a 3 ft long 2 in. pipe section inclined at 270 below the horizontal. The
area of the inlet slot is 2.5 in.2. The gas and liquid phases are metered with a gas vortex shedding
meter and a Micromotion meter, respectively, before being recombined at the loop exit. The
system is operated with air-water at atmospheric conditions.

Fig. 5 shows the effect of the operational flow rates (expressed in terms of the superficial
velocities of the gas and liquid phases in the GLCC) on d100. As can be seen in the figure, the gas
velocity does not have a significant effect on the d100 at low liquid velocities. At high liquid
velocities, however, there is an increase in d100 with increasing gas velocities. Also, as the liquid
flow rate increases, the d100 increases significantly. Note that the size of the bubbles is below 0.1
mm, as the larger bubbles are captured upstream in the vortex region.

12
0.1
Vsl = 0.05 ft/s
Vsl = 0.1 ft/s
d 100 (mm) 0.08 Vsl = 0.5 ft/s

0.06

0.04

0.02

0
0 10 20 30 40

v sg (ft/s)

Fig 5: Variation of d100 for a 3 GLCC Operated with Air-Water at Atmospheric Conditions

4
zt

2
Vsl = 0.05 ft/s
1 Vsl = 0.1 ft/s
Vsl = 0.5 ft/s
0
0 5 10 15 20 25

v t is (ft/s)

Fig 6: Effect of Tangential Velocity on Vortex Length for a 3 GLCC Operated with Air-Water
at Atmospheric Conditions

13
The effect of inlet slot tangential velocity on the length of the vortex, zt, can be seen in Fig. 6. The
result of increasing the tangential velocity is an increase in vortex length and a reduction in axial
distance available for the bubble trajectory, which can cause a reduction in the separation
efficiency.

v tis
Fig. 7 demonstrates the effect of the ratio of inlet slot tangential velocity to axial velocity, ,
vz
on d100. This ratio is significant because it determines the radial and axial distances traveled by the
bubble, which in turn affect the optimum diameter and length of the GLCC. As can be seen, the
v tis
d100 is high for low values of , but decreases sharply as the ratio is increased, above 100. For
vz
this region, due to high tangential velocity causing higher centripetal forces, and low axial
velocity resulting in larger residence time, the separation efficiency is high as manifested by the
small value of d100.

0.1
Vsl = 0.05 ft/s
0.08 Vsl = 0.1 ft/s
Vsl = 0.5 ft/s
d 100 (mm)

0.06

0.04

0.02

0
0 50 100 150 200 250

v t is /v z

Fig 7: Effect of the Ratio of Tangential Velocity to Axial Velocity on d100 for a 3 GLCC
Operated with Air-Water at Atmospheric Conditions

The effect of the tangential velocity decay on d100 is demonstrated in Fig. 8. As can be seen, for
increasing percent decay factors, the d100 increases linearly indicating lower separation efficiency.
This is more pronounced for higher operational liquid velocity conditions given by the upper line.
For our study, as indicated before, the theoretical analysis predicts a decay factor of 5.6%.

14
0.2
Vsl = 0.5 ft/s, Vsg = 5 ft/s

0.16 Vsl = 0.1 ft/s, Vsg = 10 ft/s


d100 (mm)

0.12

0.08

0.04

0
0 2 4 6 8 10
% Decay per L/D

Fig 8: Effect of Decay on d100 for a 3 GLCC Operated with Air-Water at Atmospheric
Conditions

Fig. 9 shows the bubble trajectories for a 0.4 mm bubble released at different radial locations
below the vortex region for one set of operating conditions. If the bubble is released at the
GLCC wall, Ro = Rs, it travels an axial distance Lt = 2.7 ft. As the initial radial location of the
bubble is decreased, r<Rs, the bubble requires less axial distance to be captured. For example, Lt
= 1.2 ft when the bubble starts at a distance of R0 = 0.25Rs from the centerline. The effect of
turbulence on these trajectories was examined using Eq. 18 with the positive and negative signs in
the turbulence term to represent the worst case scenarios. This analysis indicated that even for
these extreme cases, turbulence does not affect the trajectory much and can be neglected.

The most significant results of the model is presented in Fig. 10. The separation efficiency E(%)
is plotted as a function of the bubble diameter for one set of operating conditions, for 0%, 3%
and 7% decay factors. The bubble diameters for which the separation efficiency is 100% is the
d100 bubbles. As expected, larger d100 values (lower separation efficiencies) are obtained with the
higher % decay factors. For bubbles with smaller diameters than d100, the separation efficiency
drops exponentially. Note that similar plots must be generated for different operational (flow
rates) conditions.

15
0

-0.5

Length, L (ft)
-1

-1.5 Radial Location Operational


of Bubble Release: Conditions :

-2 Ro = Rs
Vsl = 0.5 ft/s
Ro = 0.9Rs Vsg = 5 ft/s
Ro = 0.75Rs P = 0 Psig
-2.5
Ro = 0.5Rs db = 0.4 mm
Ro = 0.25Rs
-3
0 0.02 0.04 0.06 0.08 0.1 0.12
Radial distance from wall, r (ft)

Fig 9: Bubble Trajectories for a 0.4mm Bubble Released at Different Radial Locations in a 3
GLCC Operated with Air-Water at Atmospheric Conditions

100
No Decay
3% Decay
80 7% Decay
% Efficiency

60

40

20

0
0 0.02 0.04 0.06 0.08 0.1 0.12

d (mm)

Fig 10: Separation Efficiency for Various Decay Factors for a 3 GLCC Operated with Air-
Water at Atmospheric Conditions

16
CFD Simulation
CFD simulations were undertaken in this study to support the mechanistic model by addressing
the following unresolved questions: 1) Does a forced vortex correctly describe the tangential
velocity profile at large axial distances from the inlet? 2) Does the bubble have to migrate all the
way to the gas core in order to be captured? 3) How should the GLCC geometry be optimized
with respect to slot dimensions, GLCC diameter and length? This can be initially studied by
looking at the effect of tangential and the axial velocities and their ratio, v t is /vz. 4) Is the effect of
turbulence predicted correctly by the mechanistic model? 5) Can CFD simulation give any insight
into the instability of the gas core? A commercially available computational fluid dynamic (CFD)
code called CFX (previously CFDS-FLOW3D) was utilized in order to study and better
understand the complex flow behavior in the GLCC configuration, and attempt to answer the
above questions.

To evaluate the accuracy of the flow field simulation, the predictions from the CFD code were
compared to experimental data (Farchi, 1985). The experimental data were carried out in a short
186 mm ID GLCC (L/D 2) operating with air and water at atmospheric conditions. A
comparison of the predicted tangential velocity profiles using a 3-D simulation showed very good
agreement with experimental data. Also, an axisymmetric (2-D) simulation of flow in the GLCC
was carried out that likewise showed very good agreement with the experimental data. Fig. 11,
for example, shows a comparison between the simulation results and the experimental data
obtained at three different probe locations below the inlet: Probe-1 is located just below the inlet,
and Probe-2 and 3 are located 100 and 200 mm below the inlet, respectively.

3
CFX Prediction Probe-1
CFX Prediction Probe-2
2.5 CFX Prediction Probe-3
Data Probe-1
2 Data Probe-2
Data Probe-3
v t ( m/s)

1.5

0.5

0
0.000 0.012 0.024 0.036 0.048 0.060 0.072 0.083

-0.5

r (m)

Fig 11: Tangential Velocity Prediction vs. Data (Farchi, 1985), ml = 2.66 kg/s, mg = 0.002 kg/s

17
Following the verification of the simulations, flow predictions were carried out for the 3 in. GLCC
described above. The simulations show that the flow in the GLCC is very complex and includes
three velocity components: the tangential velocity vt, the axial velocity vz, and the radial velocity
vr. The highest tangential velocity, vt, is observed at the inlet, as shown in Fig. 12. The high
tangential velocity dissipates quickly in the inlet region, resulting in a combined vortex, not
predicted previously. The dissipation continues in the axial direction towards the outlet, similar to
the trend shown in Fig. 11. However, this tangential velocity decay is not as drastic as the one
observed in the inlet region, and within an axial distance of L/D 1 the profile becomes that of a
forced vortex. Decay of the tangential velocity below the inlet region for the 3 in. GLCC is
shown in Fig. 12, using an axisymmetric simulation. The tangential velocity profiles, as a function
of the radial direction, in different axial locations (approximately 0,3,6,9,12, and 24 inches below
the inlet) are shown. Note that for this case, near the center of the cylinder, the tangential
velocity is nearly a constant, up to about 12 inches below the inlet. Eventually, the viscous
dissipation (boundary layer effects) influences the core region, as shown in Fig. 12, at the axial
location of 24 inches below the inlet. Furthermore, considering the dissipation away from the inlet
L
region (L/D>1) results in an approximately 7.6% decay, which compares favorably with the
D
L
5.6% decay predicted by the mechanistic model.
D

1
Location Below
0.9 Inlet (inches)
24 12
0.8
9 6
0.7 3 0

0.6
v t / v t is

0.5

0.4
0.3
0.2

0.1
0
0.00 0.01 0.02 0.03
r (m)

Fig 12: Tangential Velocity Distribution in 3 inch GLCC, vz = 0.5 ft/s, vtis = 10 ft/s

18
Fig. 13 shows the velocity vectors in a plane in front of the inlet. Secondary upward (axial) flow
is observed in center region of the GLCC, while the main axial flow goes downward with the
swirl near the wall region. The magnitude of this secondary upward flow is low as compared to
the main tangential flow, but the average axial flow is downward. This secondary flow also
decays in the axial direction from the inlet toward the outlet, depending on the magnitude of the
average axial flow velocity. Recall that in the mechanistic model a flat profile is assumed for the
axial velocity, as an approximation. This makes the mechanistic model predictions conservative
for gas carry-under.

Fig 13: Velocity Vectors in a Plane in Front of the Inlet

The axisymmetric flow velocity predictions were used to study the sensitivity of flow behavior
for different ratios of the inlet tangential velocity, v t is , to the average axial velocity, vz. These
simulations showed that the ratio of inlet tangential velocity to the average axial velocity ( v t is /vz)
has an important effect on hydrodynamic flow behavior in the GLCC. High v t is / vz ratio indicates
a strong rotational flow that enhances separation, while a low v t is / vz ratio causes a strong decay
of the tangential velocity in downward direction and causes gas carry-under.

The v t is / vz ratio has also a great affect on the capture radius. The region between the secondary
upward flow and the main downward flow (zero axial velocity) is referred to as the capture
radius, Rcap, as depicted in Fig. 14. In this region, the gas bubbles are captured and separated,
and move upward towards the gas leg. This results in a much larger capture radius than the gas

19
core radius used in the mechanistic model. This will be further investigated and incorporated in
the mechanistic model in the future.

R cap

Fig 14: Definition of Capture Radius (Rcap)

0.7

0.6

0.5
R cap /r

0.4

0.3
Location
0.2
Below Inlet
0.1 6 inches
12 inches
0
1 10 100 1000
v t is /v z

Fig 15: Variation of Capture Radius with vtis/vz

20
Predictions of the capture radius as a function of the v t is / vz ratio is shown in Fig. 15 for 6 and 12
inches below the inlet. For low v t is / vz ratios, Rcap is small, but as v t is / vz ratio increases, Rcap
increases exponentially and approaches a nearly constant value. This relationship between Rcap
and v t is / vz ratio can help define the operational envelope for gas carry-under, which is very
important for design of GLCC.

Future studies will focus on the effect of turbulence, the prediction of the gas-liquid free surface in
the vortex region and the gas core instability.

CONCLUSIONS

The Petroleum Industry has shown keen interest in utilizing GLCC separators for a wide range of
applications. A complete understanding of the hydrodynamic behavior of the flow in the GLCC is
needed to make the GLCC more predictable, reliable and viable for field applications. This paper
lays the foundation for characterizing gas carry-under and determining separation efficiency.

A bubble trajectory model has been developed to predict the course of the bubbles in the GLCC.
A model for the tangential velocity decay has also been developed, which gives results consistent
with the available empirical values used for hydrocyclones. Combination of these models allows
the determination of the separation efficiencies as a function of the bubble size. This is one of the
building blocks for a complete model for the determination of gas carry-under. Additional
experimental and theoretical work is required, on the bubble size distribution and the total amount
of entrained gas, before the goal of quantitative prediction of the overall separation efficiency can
be achieved.

NOMENCLATURE

A = cross sectional area


Cd = drag coefficient
D = diameter
d = bubble diameter
F = force
f = friction factor
g = acceleration of gravity
gc = unit conversion factor
K = resistance coefficient for fittings
k = proportionality factor
L = length
R = radius
r = radial coordinate
Re = Reynolds Number

21
V = volume
v = velocity
z = axial coordinate
zt = vortex length
P = pressure drop
= inclination angle of inlet section
= tangential velocity vector angle with respect to the horizontal
= frictional losses
= viscosity
= density
= shear stress

Subscripts and superscripts

b = bubble
c = core, centripetal
d = drag
e = effective area for capture
g = gas
in = inlet
is = inlet slot
l = liquid
r = radial
s = separator
t = tangential, turbulent, total
w = crown of vortex
z = axial
= angular
n = tangential velocity exponent
= fluctuating component of velocity

REFERENCES

Arpandi, I., Joshi A.R., Shoham, O., Shirazi, S. and Kouba, G.E., Hydrodynamics of Two-Phase
Flow in Gas-Liquid Cylindrical Cyclone Separators, SPE 30683, presented at SPE Annual
Technical Conference, Dallas, October 22-26, 1995.

Farchi, D.: A Study of Mixers and Separators for Two-Phase Flow in M. H. D. Energy
Conversion Systems M.S. thesis (in Hebrew), Ben-Gurion University, Israel, 1990.

Karamanev D.G. and Nikolov, L.N., Free Rising Spheres Do Not Obey Newtons law for Free
Seetling, Biological Faculty, Sofia University, 1421 Sofia, Bulgaria, AIChE Journal, November
1992, vol.38, No.11, pp.1843-1846.

22
Kouba, G.E., Shoham, O. and Shirazi, S., Design and Performance of Gas-Liquid Cylindrical
Cyclone Separators, Proceedings of the BHR Group 7th International Meeting on Multiphase
Flow, Cannes, France, June 7-9, 1995, pp. 307-327.

Kurokawa, J. and Ohtaki T.: Gas-Liquid Flow Characteristics and Gas Separation Efficiency in
Cyclone Separator, ASME FED vol. 225, Gas-Liquid Flows, 1995, pp. 51-56.

Kutepov, A.M., Nepomnyaahchii, E.A., Ternovskii, I.G. Pashkov, V.P. and Konovalov, G.M.:
Investigation and Calculation of the Separating Efficiency of Hydrocyclones, from J. Applied
Chemistry, USSR, English Translation, vol. 51, 1978, pp. 602-606.

Levich, V. G.: Physicochemical Hydrodynamics, Prentice-Hall, Englewood Cliffs, N. J., 1962.

Millington, B.C., and Thew, M.T.: LDA study of component velocities in air-water models of
steam-water cyclone separators, Proceeding of the 3rd International Conference on Multiphase
Flow, The Hague, The Netherlands, May 18, 1987, pp. 115-125.

Perry, J. H., Editor: Chemical Engineers Handbook, third edition, Mc Graw Hill Company,
1963.

Taitel, Y. and Dukler, A.E., A Model for Predicting Flow Regime Transition in Horizontal and
Near Horizontal Gas-Liquid Flow, AIChE J., 1976, vol.22, No.1, pp. 47-55.

Wolbert, D., Ma, B.F., Aurelle, Y. and Seareau, J.: Efficiency Estimation of Liquid-Liquid
Hydrocyclones Using Trajectory Analysis, AIChE J., vol. 41, no. 6, June 1995, pp. 1395-1402.

23

You might also like