You are on page 1of 8

Available online at www.sciencedirect.

com

Journal of Ocean Engineering and Science 5 (2020) 261–268


www.elsevier.com/locate/joes

Evaluation of flow characteristics in an onshore horizontal separator using


computational fluid dynamics
Tathagata Acharya∗, Lucio Casimiro
Department of Physics and Engineering, California State University, Bakersfield, 9001 Stockdale Highway, Bakersfield, CA 93311, United States
Received 8 August 2019; received in revised form 25 November 2019; accepted 26 November 2019
Available online 10 January 2020

Abstract
In the petrochemical industry, separation of oil from water is a very important process. Wells produce mixtures of gas, oil, and water
which undergo a primary stage of separation inside horizontal gravity separators. The performance of these vessels is evaluated by measuring
mean residence time (MRT) and residence time distribution (RTD). Although many researchers studied flow characteristics in horizontal
separators, limited number of articles exist that discuss separator MRT and RTD with varying water-cuts. In this article, the authors study an
experiment using a horizontal gravity separator by previous researchers and perform computational fluid dynamics (CFD) simulations on the
same geometry under similar conditions. The simulation results show qualitative agreement with the experiments by previous researchers.
As shown by experiments before, CFD results showed that MRT of the organic phase increased with increase in water-cut. In addition, the
RTD characteristics show very similar trends between CFD and experimental results.
© 2020 Shanghai Jiaotong University. Published by Elsevier B.V.
This is an open access article under the CC BY-NC-ND license. (http://creativecommons.org/licenses/by-nc-nd/4.0/)

Keywords: Multiphase separators; Computational fluid dynamics; Mean residence time; Residence time distribution.

1. Introduction vertical separators, horizontal separators are typically used


for produced fluids with lower gas-oil ratio (GOR) [1–4].
Wells produce mixtures of oil, water, and gasses. Dur- Therefore, horizontal separators are usually large vessels that
ing subsea operations, the produced fluids from different are required to handle large volumes of liquids. Previous re-
wells are collected in subsea manifolds. The subsea flowlines searchers have also shown that horizontal separators are more
carry these fluids to offshore platforms, from where these economical and enable better processing operations when
are transported onshore using trunklines. The produced mix- emulsions and foam are present. However, these separators
tures are then introduced to multiphase separators, which are are not as effective with sand or surge processing [4]. The
vessels separating the various phases in produced mixtures. cyclone separator is a compact, and low-cost separator that
Fig. 1 shows a schematic of offshore production with onshore consists of a vertical pipe with tangential inlets and outlets
multiphase separators. for gas and liquid. With tangential flow from the inlet to the
Based on orientation, multiphase separators can be clas- body of the separator, the flow swirls with enough tangential
sified as horizontal, vertical, and cyclone separators. Vertical velocity causing centripetal forces in the entrained gas. The
separators occupy lesser plot area and are easier to transport combined effect of centripetal forces and gravity forces push
than horizontal separators. Therefore, vertical separators the liquids radially outward making them drop by gravity to-
are used when produced fluids have smaller liquid loads, wards the liquid exit. The gas moves radially inward and then
or when the availability of space is limited. Compared to moves upwards and is collected at the gas exit. Multiphase
separators can either be two-phase or three-phase based on
∗ Corresponding author.
their functionality. While three phase separators separate gas,
E-mail addresses: tacharya@csub.edu (T. Acharya), lcasimiro@csub.edu oil, and water, two phase separators separate oil from water
(L. Casimiro). or oil from gas. The petroleum industry incurs huge costs
https://doi.org/10.1016/j.joes.2019.11.005
2468-0133/© 2020 Shanghai Jiaotong University. Published by Elsevier B.V. This is an open access article under the CC BY-NC-ND license.
(http://creativecommons.org/licenses/by-nc-nd/4.0/)
262 T. Acharya and L. Casimiro / Journal of Ocean Engineering and Science 5 (2020) 261–268

platform onshore manifold


Nomenclature trunkline
onshore separator1

MRT mean residence time, s


c(t) concentration of a given phase as a function of
onshore separator2
time
t time, s
RTD residence time distribution
onshore separator3
MRTA mean residence time evaluated using active vol- flowline
ume occupied by a phase, s
VA active vessel volume occupied through with
each phase flow, m3 subsea manifold
∀˙ volume flowrate, m3 /s
subsea wells
ρ density, kg/m3
u fluid velocity in horizontal direction, m/s
v fluid velocity in vertical direction, m/s Fig. 1. Subsea petroleum drilling.
μ dynamic viscosity, Pa-s
α volume fraction of one phase in a two-phase
mixture density ratio between gasses and liquids in gas-liquid systems
β volume fraction of the other phase in the two- is in the range of 0.001–0.1, it is in the range of 0.7–1.1 in
phase mixture liquid–liquid systems. Also, viscosity of the oils in liquid-
vr oil rising velocity in water, m/s liquid systems vary a lot. Therefore, it may be more difficult
ρp density of the discrete phase, kg/m3 to predict flow characteristics of liquid–liquid systems than
ρf density of the continuous phase, kg/m3 gas-liquid systems.
g acceleration due to gravity, m/s2 Phase separation in a horizontal three-phase separator is
R radius of the oil droplet in water, m accomplished in three zones [15]:
lw thickness of the water layer, m
(a) Primary zone: In this zone, an inlet diverter is used to
separate bulk of the liquid phase. The abrupt changes
towards transporting hydrocarbons. Since phase separation in flow velocity and direction cause the largest liquid
is done before hydrocarbons are transported, optimizing the droplets to impinge on the diverter and drop by gravity.
sizes of these vessels is extremely important to the industry. (b) Gravity separation zone: In this zone, vapor and liquid
Many articles exist which detail procedures to design mul- phases flow with low velocities and little turbulence.
tiphase separators. A lot of these procedures are derived This causes gravity separation of fine droplets out of
through practice and are based on thumb rules. For exam- the gas phase. This zone also contains a liquid collec-
ple, several heuristic guidelines were proposed for separator tion zone where entrained gas bubbles, and other liquid
aspect ratio. Aspect ratio is defined as the ratio of length or droplets join their phases due to buoyancy and gravity.
height of the separator to the diameter of the separator. Smith (c) Mist elimination zone: This is the zone where very fine
suggested that the aspect ratio should be in the range of 1.0– droplets are separated by passing the gas stream through
9.0. It was further suggested that if gas flowrates determined demister. This zone may have coalescing plates, vanes,
the size of the vessel, the preferred aspect ratio range should and wire mesh pad to provide an impingement surface
be 2.0–6.0 for horizontal separators and it should be 2.0–3.0 for very fine droplets, such that they can coalesce and
for vertical separators. If liquid flowrates determined the ves- form larger droplets and can be separated out of gas
sel size, then the aspect ratio range should be 2.0–6.0 [5]. stream by gravity.
Walas suggested that aspect ratio should be in the range of
2.0–5.0 based on separator operating pressure [6]. Grødal and
Realff concluded that the upper limit of aspect ratio is a strong Fig. 2 shows the construction of a typical horizontal
function of the physical properties of the fluid which are dif- multiphase separator.
ferent for different oilfields [7]. Watkins developed guidelines Since multiphase separators may need to handle streams
that suggested increasing the length of the separator to make from several different wells, it requires them to be very flex-
the aspect ratio equal to 3 if it was less than 3. Also, it was ible to handle different mixture flowrates. Intermittent flow
suggested that a horizontal separator should be chosen if the patterns like slug flows can cause interface control problems.
aspect ratio of a vertical separator was more than 5 [8]. Also, solid deposits may create blockage within the separator.
Phase separation in separators is a relatively under- The performance of multiphase separators is often measured
researched area. Although some literatures exist on gas–liquid based on the mean residence time (MRT) of the oil phase
separation [9–14], not much data is available on separation inside the separator. Mean residence time is defined as the
between immiscible liquids such as oil and water. While the amount of time a given phase stays inside the separator and
T. Acharya and L. Casimiro / Journal of Ocean Engineering and Science 5 (2020) 261–268 263

Flow inlet
Vapor outlet
Continuous progress in the field of numerical methods has
led to computational fluid dynamics (CFD) being an alterna-
tive tool gaining more confidence among the industrial com-
diverter Mesh pad munity [23–26]. CFD is more flexible and less expensive
than fluid flow experiments that require costly instrumenta-
Liquid level Weir tion. It is also more universal than semi-empirical models
Oil-water interface used in designing of separators. Previous research in this
area generally involved model development accounting for
droplet size distribution in addition to coalescence and break-
up [27–30]. Fewer articles focused on assessing effects of sep-
Water outlet Oil outlet arator internals on separator performance [31–33]. However,
there is a lack of published CFD studies that directly eval-
Fig. 2. Horizontal separator components. uate phase MRT based on RTD. Such studies are extremely
critical towards understanding flow characteristics in multi-
is shown by Eq. (1) [16]: phase separators and more accurate assessment of separator
∞ performance.
t c (t )dt
MRT = 0∞ , (1) While very few articles exist that evaluate MRT based on
0 c (t )dt RTD experimentally [16,22], seemingly none of the previ-
where MRT is the mean residence time, c(t) is the concen- ously published CFD work report RTD based MRT evalua-
tration at the outlet, and t is time in seconds. With increase tions in oil-water mixtures with varying volume-fractions of
in MRT, a given phase has more time to disengage from the water. In the present work, the authors perform CFD simu-
other phase. lations to study liquid-liquid flow characteristics in a simple
Residence time distribution (RTD) is defined as the fraction horizontal multiphase separator. The authors use CFD results
of elements leaving the separator between times t and t + t. to obtain RTD and MRT as functions of varying water-cuts.
Following is the equation for RTD [16]: These results are further compared with experiments using a
similar geometry by previous researchers.
c (t )
RT D (t ) =  ∞ . (2)
0 c (t )dt 2. Simulation geometry and methods
Hence
 ∞ The geometry contains an inlet, two outlets, a weir, and
RT D (t ) = 1. (3) a porous plate but doesn’t accommodate components such
0
as the demister, mesh-pad, and the diverter. The geometry
Eq. (3) suggests that over a long period of time, cumulative
is shown in Fig. 3. A weir separates the water outlet from
residence time distribution will be equal to unity. Therefore,
the oil outlet. Two different heights of weir are used for the
while evaluating MRT, the upper limit of the integral is chosen
purpose of simulations. From CFD results, the authors report
such that cumulative RTD tends to unity.
MRT and RTD with varying water-cuts.
Through the past several decades, researchers have had
Two phase simulations are performed with mixtures of wa-
varying opinions about optimal MRT for multiphase separa-
ter and kerosene in a horizontal separator. The separator is
tors [3,17–21]. On field operators usually measure MRT using
0.6 m in diameter and 2.5 m in length. The simulations are
an alternate formula shown by Eq. (4):
performed using two different weir heights: 0.22 m and 0.3 m.
VA The separator inlet and outlet diameters are 0.05 m each. The
MRTA = , (4)
∀˙ separator is classified into three zones – an inlet zone with
where MRTA is mean residence time evaluated using active the flow inlet, a porous zone, and the separation/outlet zone
vessel volume, VA is the active vessel volume involving each with the two outlets. The length of the outlet zone of the sep-
phase, and ∀˙ is the volumetric flowrate. The volume occu- arator is 2.18 m, and the weir is placed halfway between the
pied by each phase is calculated assuming that the height of water and oil outlets. The height of the inlet and the two out-
oil-water interface can be extrapolated from the weir back to lets is 0.3 m. The porous zone is modeled as a 0.20 m thick
the separator inlet. This method may involve substantial er- perforated plate with a viscous resistance of 5 × 108 m−2 .
rors associated with oil-water interface height obtained using Simulations are performed with four different volume frac-
level controllers [16,22]. Therefore, evaluation of MRT using tions of water (water-cuts): 10%, 21%, 32%, and 57%. In the
RTD information, as shown by Eq. (1) is desirable for better mixture of kerosene and water, for each water-cut, the mass
accuracy. flowrate of water is kept constant at 4 kg/s. For each water-
On field RTD and corresponding MRT measurements are cut, transient simulations are performed through a flowtime
difficult because they may require complicated instrumenta- of 40 s.
tion and flow visualization. This may sometimes be virtually 2-dimensional transient, multiphase CFD simulations are
impossible because field vessels operate under pressure and performed on the horizontal separator using ANSYS Flu-
are constructed of steel [22]. ent 18.2. The code uses finite volume method to solve
264 T. Acharya and L. Casimiro / Journal of Ocean Engineering and Science 5 (2020) 261–268

Fig. 3. Simulation geometry with the small weir.

Navier Stokes equations in two dimensions as shown by


Eqs. (5)–(7).
∂ρ ∂ (ρu ) ∂ (ρv )
+ + = 0, (5)
∂t ∂x ∂y
 
∂u ∂u ∂u 1 ∂P μ ∂ 2u ∂ 2u
+u +v =− + gx + + 2 , (6)
∂t ∂x ∂y ρ ∂x ρ ∂ x2 ∂y
 2 
∂v ∂v ∂v 1 ∂P μ ∂ v ∂ 2v
+u +v =− + gy + + 2 , (7)
∂t ∂x ∂y ρ ∂y ρ ∂ x2 ∂y
where ρ is the density of the fluid, u is fluid velocity along the
horizontal direction, v is the fluid velocity along the vertical
direction, P is the static pressure, and μ is the dynamic vis-
cosity of the fluid. Eq. (5) shows the continuity or mass con-
servation equation, and Eqs. (6) and (7) show the momentum
conservation equations in x and y directions within the flow Fig. 4. Flow characteristics at time t = 5 s.
domain. These equations may be derived using control vol-
ume theory. Since the problem does not involve heat transfer,
the energy equation is not solved. The code uses the Eulerian where α is the volume fraction of a given phase. Since the
multiphase model. In this model, a single pressure is shared multiphase mixture involves two fluids only, the volume frac-
by all phases, and the momentum and continuity equations tion of the second phase, β is given by Eq. (11):
are solved for each phase. Volume fraction for both phases β = 1 − α. (11)
(kerosene and water) are used as inputs. Eqs. (8)–(10) show
The k-ε turbulence model is used since flow through the
the modified version of continuity and momentum equations
inlet is turbulent in nature. The geometry is meshed us-
solved for each phase using the Eulerian multiphase model
ing ANSYS meshing tool. Quad mesh is used, and mesh
[34]:
independence is obtained with 240,552 nodes.
∂ ∂ ∂
(αρ) + (αρu ) + (αρv ) = 0, (8)
∂t ∂x ∂y 3. Results and discussion

∂ ∂ ∂ The separator is empty at time t = 0. Figs. 4 and 5 show


(αρu ) + u (αρu ) + v (αρu )
∂t ∂x ∂y the flow inside the separator after 5 s, and 40 s from start
    with 21% water-cut and using the high weir. After 5 s from
∂P ∂ ∂u ∂ ∂u
= −α + αρgx + αμ + αμ , (9) start, the liquids flow over the weir and after 40 s, liquids
∂x ∂x ∂x ∂y ∂y
exit from both oil and water outlets.
Figs. 6, and 7 show the velocity profiles of kerosene after
∂ ∂ ∂
(αρv ) + u (αρv ) + v (αρv ) 5 s, and 40 s from start respectively. After 40 s from start, the
∂t ∂x ∂y
    kerosene velocity profile shows higher kerosene velocities in
∂P ∂ ∂v ∂ ∂v upper regions of the mixture of kerosene and water. This sug-
= −α + αρgy + αμ + αμ , (10)
∂y ∂x ∂x ∂y ∂y gests gravity separation in the region with water falling below
T. Acharya and L. Casimiro / Journal of Ocean Engineering and Science 5 (2020) 261–268 265

Fig. 5. Flow characteristics at time t = 40 s. Fig. 7. Kerosene velocity profile at time t = 40 s.

Fig. 6. Kerosene velocity profile at time t = 5 s.


Fig. 8. Kerosene velocity vectors across the porous zone.

kerosene. Fig. 8 shows velocity vectors of kerosene, enter-


ing and exiting the perforated plate. The velocity of kerosene RTD versus time - low weir
phase substantially reduces as it passes through the perforated 0.45

plate. The perforated plate is closer to the inlet and provides 0.4
10 % water-cut
0.35
resistance to flow. It is useful towards increasing MRT. 21 % water-cut
0.3
The RTD and MRT are evaluated as functions of water-cut. 0.25
32 % water-cut
RTD(t)

Eq. (2) is used to calculate RTD at each water-cut. Figs. 9 and 0.2
10 show RTD versus time with the low weir and high weir 0.15 57 % water-cut

respectively. For both configurations, RTD reaches its peak 0.1

value very close to the start. With both weir configurations, 0.05
0
the RTD peak is the highest with 10% water-cut. The RTD 0 5 10 15 20 25 30 35 40 45
-0.05
peak reduces as water-cut is increased to 21%. This suggests time (seconds)
that with larger volume fraction of water in the mixture, lesser
amount of kerosene leaves the separator from start through Fig. 9. RTD versus time with low weir.
a given time corresponding to the peak. However, with the
low-weir configuration, the RTD peak increases marginally
from 21% water-cut to 32% water-cut. With both configura- calculations. At 57% water-cut smaller secondary RTD peaks
tions, the RTD peaks marginally shift towards the right with are observed with the high weir. The presence of secondary
increasing water-cut. This indicates that as amount of wa- peaks generally indicates secondary flows within the separator
ter in the mixture increases, kerosene takes longer to leave due to considerable differences between the individual phase
the separator. These results could further be verified by MRT velocities [16].
266 T. Acharya and L. Casimiro / Journal of Ocean Engineering and Science 5 (2020) 261–268

RTD versus time - high weir


0.5
0.45 10 % water-cut
0.4
0.35 21 % watercut
0.3 32 % watercut
RTD(t)

0.25
57 % watercut
0.2
0.15
0.1
0.05
0
-0.05 0 5 10 15 20 25 30 35 40 45
time (seconds)

Fig. 10. RTD versus time with high weir.


Fig. 12. Kerosene droplet rising velocity in water.

MRT versus watercut Using CFD results, the oil droplet rising velocity is calcu-
8 lated using Eq. (13):
7 lw
vr = , (13)
Mean Residence Time (MRT)

6 MRT
5
where lw is the water layer thickness, and MRT is mean res-
idence time [29]. Kerosene rising velocity in water obtained
4 using CFD results is compared against the rising velocity
3 obtained using Stokes Law assuming a droplet diameter of
High Weir 500 μm. Fig. 12 shows a comparison of kerosene rising ve-
2
Low Weir locity in water obtained using both methods.
1 Both methods show that kerosene rising velocity in wa-
0
ter increases with increase in water-cut. However, velocities
0 10 20 30 40 50 60 obtained from CFD results are slightly larger than the veloc-
percentage water-cut ities obtained from Stokes Law. Therefore, the classic design
methods maybe somewhat more conservative when estimating
Fig. 11. MRT versus water-cut. rising velocity of kerosene droplets. This observation agrees
with results published by previous researchers [29].

Eq. (1) is used to calculate MRT at each water-cut, for 4. Experimental validation
both weir configurations. Fig. 11 shows that MRT of the
organic phase (kerosene) increases with increase in water-cut The authors compared their numerical results against very
for both weir configurations. With increase in weir height, similar experiments performed by previous researchers [16].
MRT increases for each water-cut. Therefore, as water-cut in- However, since some pieces of information about the experi-
creases, the organic phase is retained within the separator for mental geometry were unavailable, only a qualitative compari-
longer duration. With increase in retention time, the organic son between the results was possible. The details are provided
phase has more time to disengage from the aqueous phase. in the following sections.
When the weir height is increased, the organic phase is also Simmons et al. performed pilot-scale experiments using a
retained for durations longer than the lower weir height. horizontal separator and measured MRT and RTD for vary-
During phase separation oil rising velocity in water can be ing water-cuts with two different weir-heights [16]. Like the
obtained using Stokes law, as shown by Eq. (12): simulation geometry, the separator used by them also had a
  diameter of 0.6 m and a length of 2.5 m and used two differ-
2 ρ p− ρ f ent weir heights: 0.22 m and 0.3 m. A two-phase mixture of
vr = gR2 , (12)
9 μ kerosene and water was used. To measure RTD, they injected
tracer dyes that were soluble in only one phase at a time.
where vr is the oil droplet rising velocity, ρ p is the density of MRT and RTD were measured using Eqs. (1) and (2).
the discrete phase, ρ f is the density of the continuous phase, They showed that with both weir configurations, MRT of
μ is the dynamic viscosity of oil water emulsion, g is the the organic phase (kerosene) increased with increase in water-
acceleration due to gravity, and R is the radius of oil droplet cut. Also, with the high weir, MRT increased with all three
in water. Previous studies show that diameters of freely ris- water-cuts than the low weir. Simmons et al. measured RTD
ing stable untreated oil droplets range from 50 μm to 1 mm as a function of water-cut and showed that RTD peaks were
[35,36] shifted towards the right with increase in water-cut. The RTD
T. Acharya and L. Casimiro / Journal of Ocean Engineering and Science 5 (2020) 261–268 267

Table 1 area focus on droplet size distribution and evaluating separa-


Comparison of results with Simmons et al.’s experiments [16]. tor internals, the authors report CFD results evaluating MRT
Sr No. Numerical results that agree with experiments and RTD as functions of water-cut. The authors also show
1 Kerosene MRT increases with increase in water-cut CFD can be used to obtain valuable information about liquid-
2 Higher MRT values are observed with the high weir than with liquid flow characteristics in a horizontal separator avoiding
the low weir. expensive instrumentation that may be required to measure
3 RTD peaks occur closer to the start MRT and RTD. The authors believe these results provide
4 RTD peaks generally shift towards the right with increase in
water-cut
valuable pieces of information to future researchers involved
5 Secondary RTD peaks are observed at larger water-cuts such as with designing of liquid-liquid multiphase separators.
at 57%
6. Future work
peak increased with increase in water-cut from 36% to 49% Current availability of computational resources precludes
with the low weir configuration. All these experimental re- full-scale 3-dimensional CFD simulations. However, in near
sults are in direct agreement with the CFD results by the future, the authors intend to run 3-dimensional simulations
authors. Table 1 summarizes these results between CFD and to study the effect of side baffles in the geometry. Also,
the experiment by Simmons et al. additional simulations with different hydrocarbons will be
There are a few dissimilarities between the geometry simu- performed to study flow characteristics.
lated and the geometry used for experimentation by Simmons
et al., which are as follows:
Declaration of Competing Interest
Simmons et al.’s report did not mention the inlet and outlet
diameters and details about the perforated region. In addition,
The authors declare that there are no conflicts of interest.
the exact location of the weir, the inlet and the two outlets
are not reported by them. These pieces of information were
Acknowledgment
assumed by the authors for their simulations. The experimen-
tal geometry used side baffles to simulate dead zones within
The authors would like to thank CSUB undergraduate stu-
the separator. Since the authors performed 2D simulations,
dents Brooke Riehl and Nicholas Bustamante for their help
side baffles could not be used. However, despite these minor
with literature survey.
differences, the CFD results by authors confirm the general
trends obtained from experiments by the previous researchers.
References

5. Conclusion [1] Gas Processors Suppliers Association, GPSA Engineering Data Book,
Volume 1, eleventh ed., Gas Processors Association, Tulsa, 1998.
Computational fluid dynamics (CFD) simulations are per- [2] G. Skelton, P.A. Stockil (Ed.), British Petroleum Company Limited,
1977.
formed to assess kerosene-water flow characteristics inside [3] F. Evans, Equipment Design Handbook for Refineries and Chemical
a horizontal multiphase separator. Kerosene mean residence Plants, Volume 2, 1974 Gulf Houston, ISBN: 978-0-87-201255-4.
time (MRT) is evaluated based of residence time distribution [4] K. Arnold, M. Stewart, Surface Production Operations, third ed., Else-
as this approach leads to more accurate results. Simulations vier, New York, 2008 ISBN: 978-0-7506-7853-7.
[5] H.V. Smith, Petroleum Engineering Handbook, Society of Petroleum
show that in a two-phase mixture of immiscible liquids such
Engineering, Richardson, TX, 1987.
as water and kerosene, MRT of the organic phase (kerosene) [6] S.M. Walas, Chemical Process Equipment Selection and Design, But-
increases with increase in water-cut. Therefore, separator per- terworth-Heinemann, Houston, 1990, pp. 713–715.
formance improves with increase in water-cut. Also, between [7] E.O. Grødal, M.J. Realff, Optimal Design of Two- and Three-Phase
the two weir heights used, higher MRT values are obtained Separators: A Mathematical Programming Formulation, Society of
Petroleum Engineering, 1999, pp. 1–16. 56645.
with the higher weir for each water-cut. This is intuitive
[8] R.N. Watkins, Hydrocarbon Process. 46 (11) (1967) 253–256.
as with the higher weir, the liquids are retained within the [9] A. Hallanger, F. Soenstaboe, T. Knutsen, in: SPE Annual Technical
separator vessel for longer durations. On the RTD versus time Conference and Exhibition, Denver, CO, 1996, pp. 695–706, doi:10.
plot, RTD peaks shift towards the right with increase in water- 2118/36644-MS.
cut. Since RTD is defined as the fraction of a given phase that [10] T.T. Le, S.N. Ngo, Y. Lim II., C.K. Park, B.-.D. Lee, B.-G. Kim, D.-
H. Lim, J. Petrol. Sci. Eng. 171 (2018) 731–747, doi:10.1016/j.petrol.
stays within the separator between times t and t + t, this
2018.08.001.
suggests that with less quantity of water, larger volume frac- [11] P. Yu, S. Liu, Y. Wang, W. Lin, Z. Xiao, C. Wang, Procedia Eng. 31
tions of kerosene leave the separator closer to the start. Also, (2012) 145–149, doi:10.1016/j.proeng.2012.01.1004.
at larger water-cuts such as 57%, secondary smaller RTD [12] A. Ghaffarkhah, M. Shahrabi, M. Moraveji, H. Eslami, Egypt. J. Petrol.
peaks are observed. This suggests secondary flow behavior 26 (2) (2017) 413–420.
within the separator which may rise due to large differences [13] N. Kharoua, L. Khezzar, H. Saadawi, in: ASME 2013 Fluids Engineer-
ing Division Summer Meeting, 2013 V01CT17A013-V01CT17A013.
in velocities between the phases. The simulation results by [14] E. Hansen, ASME-PUBLICATIONS-PVP 431 (2001) 23–30.
the authors agree with previous experimental observations by [15] A.P. Laleh, W.Y. Svrcek, W.D. Monnery, Can. J. Chem. Eng. 90 (2012)
Simmons et al.. While almost all previous CFD articles in this 1547–1560.
268 T. Acharya and L. Casimiro / Journal of Ocean Engineering and Science 5 (2020) 261–268

[16] M.J.H. Simmons, J.A. Wilson, B.J. Azzopardi, Chem. Eng. Res. Des. [27] M. Abdulkadir, V.H. Perez, World Acad. Sci. Eng. Technol. 61 (2010)
80 (5) (2002) 471–481. 35–43.
[17] J.R. Couper, W.R. Penney, J.R. Fair, S.M. Walas, Chemical Process [28] N. Kharoua, L. Khezzar, H. Saadawi, in: Proc. ASME FEDSM, 2013,
Equipment- Selection and Design, Gulf Professional Publishing, 2012 p. 16321.
ISBN: 978-0-12-396959-0. [29] B.A. Grimes, J. Dispers. Sci. Technol. 33 (2012) 578–590.
[18] A. Gerunda, Chem. Eng. 88 (9) (1981) 81–84. [30] B.A. Grimsea, C.A. Doraob, N.V.D.T. Opedala, I. Kralovaa,
[19] W.Y. Svrcek, W.D. Monnery, Chem. Eng. Prog. 89 (10) (1993) 53–60. G.H. Sørlande, J. Sjöbloma, J. Dispers. Sci. Technol. 33 (2012)
[20] R.K. Sinnott, Chemical Engineering Design, Elsevier, 2014 ISBN: 591–598.
978-1-48-329470-4. [31] D. Wilkinson, B. Waldie, M.I. Mohamed Nor, H. Yen Lee, Chem. Eng.
[21] K. Arnold, M. Stewart, Surface Production Operations, third ed., Else- J. 77 (2000) 221–226.
vier, New York, 2008. [32] A.P. Laleh, W.Y. Svrcek, W.D. Monnery, Oil Gas Facil. 1 (6) (2012)
[22] M.J.H. Simmons, E. Komonibo, B.J. Azzopardi, D.R. Dick, Chem. Eng. 57–68.
Res. Des. 82 (10) (2004) 1383–1390. [33] A.P. Laleh, W.Y. Svrcek, W.D. Monnery, Oil Gas Facil. 2 (1) (2013)
[23] O. Abu Arqub, M. Al-Smadi, Numer. Methods Partial Differ. Equ. 34 52–59.
(5) (2018) 1577–1597. [34] Fluent, A.N.S.Y.S., 2013, ANSYS Fluent Theory Guide 15.0. ANSYS,
[24] O. Abu Arqub, Numer. Methods Partial Differ. Equ. 34 (5) (2018) Canonsburg, PA
1759–1780. [35] S. Hu, R.C. Kinter, AIChE J. 1 (1) (1955) 42–48.
[25] O. Abu Arqub, Fundam. Inform. 166 (2) (2019) 87–110. [36] P.J. Brandvik, Ø. Johansen, F. Leirvik, U. Farooq, P.S. Daling, Mar.
[26] O. Abu Arqub, Fundam. Inform. 166 (2) (2019) 111–137. Pollut. Bull. 73 (1) (2013) 319–326.

You might also like