Вы находитесь на странице: 1из 8

02638762/04/$30.00+0.

00
# 2004 Institution of Chemical Engineers
Trans IChemE, Part A, October 2004
Chemical Engineering Research and Design, 82(A10): 13831390

RESIDENCE TIME DISTRIBUTIONS AND FLOW BEHAVIOUR


WITHIN PRIMARY CRUDE OIL WATER SEPARATORS
TREATING WELL-HEAD FLUIDS
M. J. H. SIMMONS1 , E. KOMONIBO2, B. J. AZZOPARDI2 and D. R. DICK3
1
Department of Chemical Engineering, University of Birmingham, Edgbaston, Birmingham, UK
School of Chemical, Environmental and Mining Engineering, University of Nottingham, Nottingham, UK
3
BP Exploration, Sunbury on Thames, UK

ithin primary crude oil separators used by the oil industries, the residence time distribution of both organic and aqueous phases has been obtained for the purpose of
flow diagnostics. This paper describes the application of the Alternative Path
Model developed by Simmons et al. (2002) to give a quantitative description of the hydrodynamics and mixing within several field separators. Parameters developed from the model are
used to describe the degree of mixing within the vessels. The model shows that vessel performance is affected by the internal configuration (flow smoothing baffles and separation
plates) and the primary separation duty (gas liquid or oil water). The presence of baffling
is shown to reduce the turbulence within the flow for oil water separation, but less so for
gas oil separation, which had the overall effect of increasing mixing levels, perhaps due to
the buoyancy of the fast rising gas bubbles. The presence of secondary peaks on some of
the measured residence time distributions indicates the presence of secondary flows within
the main body of the separators. This was most noticeable when the differential velocity
between the oil and water phases was high.
Keywords: multiphase flow; liquid liquid mixing; residence time distribution; crude oil water
separators; gravity settling.

INTRODUCTION
Gravity separation of the mixture of fluids produced from
petroleum reservoirs is used to achieve a primary split
between the gas, oil, and water phases. This operation is
necessary since it is the usual practice to separate the
phases before pumping to downstream process facilities.
This is done in order to remove the water phase and also
to prevent operational difficulties, such as the presence of
slug flow in the pipelines, which may occur. Since the
volume flow rates of the fluids produced are very large,
the separation process is usually performed in a train of
horizontal cylindrical vessels.
Recent trends in the design of offshore platforms and fluid
separation equipment are aimed at reducing costs by saving
on space and weight, which creates considerable motivation
for the development of methods to improve separation efficiency. To achieve this, it is necessary to accelerate the
separation process. The major factors controlling the separation of the phases are the settling and coalescence of drops

Correspondence to: Dr M.J.H. Simmons, Department of Chemical Engineering, University of Birmingham, Edgbaston, Birmingham, B1S 2TT, UK.
E-mail: m.j.simmons@bham.ac.uk

(Gerunda, 1981). Both of these processes can be accelerated


by augmenting the gravitational force, which has led to
development of centrifugal devices, such as hydrocyclones
and also electrostatic devices (Bailes and Larkai, 1981,
1982). However, these methods do not have sufficient flexibility because adverse factors affect the operation of the separation train. These factors include fluctuations in the flow
rate, changes in the water cut (the volume fraction of
water) and the solids content, the flow pattern at the vessel
inlet and variations in the physical properties of each
phase. For these reasons, the degree of separation obtained
between the phases, particularly for oil and water, can be
poor and high cross-entrainments are observed.
The most recent approaches have been directed towards
improving the efficacy of gravity settlers. Effort has been
focused upon the modification of the vessel internals
either by introducing structured packing or perforated
baffle plates to promote coalescence and smooth the flow
respectively (Meon et al., 1993; Nilsen and Davies, 1996;
Rowley and Davies, 1988; Simmons et al., 2002;
Wilkinson et al., 2000).
Despite the many improvements made in the application
of computer modelling techniques such as computational
fluid dynamics (CFD) for multiphase flows (Hansen et al.,

1383

1384

SIMMONS et al.

1991; Mohamad Nor and Wilkinson, 1998; Rashad et al.,


1997; Wilkinson and Waldie, 1994; Wilkinson et al. 2000)
these cannot be used to predict the flow field within the
separators a priori due to the highly complex multiphase
hydrodynamics present. A particular problem experienced
in model validation is the varying physical properties of
the phases in the separators due to the presence of contaminants. It is therefore necessary to complement any modelling
with flow visualization; however, this is impossible for field
vessels since they operate under pressure and are constructed
from steel. An alternative approach is to deduce the flow
behaviour by obtaining residence time distributions (RTD)
for each phase. This approach has been used to yield qualitative information on field vessels (Dick, 1997). Simmons
et al. (2002) recently measured RTDs on a one-fifth scale
model (pilot-scale) of a primary separator operated by BP
Oil on the Forties Oil Production Facility in the North Sea.
This work introduced the Alternative Path Model (APM)
as a means of developing quantitative parameters to describe
the level of mixing.
In this paper, experimental measurements of RTD made
for British Petroleum for a selection of field separators are
presented (Dick, 1997). The modelling approach described
by Simmons et al. (2002) is applied to obtain parameters
that are used to determine the relative levels of turbulence
and mixing. This information is used to assess the performance of each separator with respect to the internal configurations and the operating parameters.
EXPERIMENTAL
Equipment
The RTD data was provided by British Petroleum (BP)
for several field primary separators. These separators are
Ula, Norway; Kinneil, UK; Milne Point, Alaska and
Magnus, UK (Dick, 1997). Schematics of the vessels are
given in Figure 1, which shows that the internals employed
in each vessel vary considerably in sophistication. The Ula
and Magnus vessels might be described as the most basic
type employed in this study (Figure 1a). They contain
a momentum breaker at the inlet to attempt to reduce the
turbulence caused by the jet of liquid impacting on the
free surface of liquid within the vessel. A baffle is used
to separate the oil and water phases and the oil and water
outlets are equipped with vortex breakers to prevent
entrainment of other phases due to vortices. A demister
pad is installed at the gas outlet to knock out any fine particles of liquid. These vessels are used primarily for the
three-phase disengagement of gas, oil, and water.
The Milne Point vessel (Figure 1b) is somewhat more
complex in configuration and also used for three-phase separation. The risk of slugging, due to the configuration of the
phases in the inlet pipework, has heavily influenced the
design of the inlet section, where the feed passes down
an expanding pipe through the length of the vessel in
order to dampen out any fluctuations in the flow. The
vessel is equipped with a perforated baffle downstream of
the vessel inlet to attempt to smooth the flow through the
main body of the vessel. Vane packs are installed to attempt
to accelerate droplet coalescence. The outlet design is similar to that used applied to the Ula and Magnus vessels
(Figure 1a) except that an elevated vortex breaker is used

Figure 1. Schematic diagrams of the field separators.

instead of a baffle to separate the oil phase from the


water phase.
The Kinneil vessel, shown in Figure 1c, makes use of
a perforated baffle at the inlet combined with a cyclonic
inlet device that extends beneath the liquid surface in the
vessel. The purpose of the perforated baffle is to smooth
the flow through the vessel, as mentioned previously,
while the cyclonic inlet device is present to assist in the
knock out of the gas phase. Further smoothing baffles are
installed along the length of the vessel to minimize the
possibility of formation of recirculation zones. This
vessel differs from the others in that the main duty is the
separation of gas and liquid phases and only very small
amounts of water are present in the feed. All the other
vessels are used primarily for oil water separations.
Measurement Techniques
The significance of the distribution of residence times
in continuous flowing systems was first presented by

Trans IChemE, Part A, Chemical Engineering Research and Design, 2004, 82(A10): 13831390

RESIDENCE TIME DISTRIBUTIONS IN OIL WATER SEPARATORS


Danckwerts (1953) and a thorough explanation of the
experimental methodologies required to obtain RTDs is
given in Levenspiel (1999). Critical to the determination
of the RTD is selection of an appropriate tracer. This is
usually injected at the inlet of the flow system and by monitoring of the concentration of the tracer at the system outlets, the RTD may be obtained. Ingersoll (1951) discussed
the use of appropriate chemical tracers to investigate the
hydraulic performance of separators. The measurement of
the RTDs presented in this study involved the use of a
radioactive tracer, Bromine-82, which was introduced
into the inlet process stream as an instantaneous pulse
(Dick, 1997). Oil-soluble and water-soluble tracers were
used to enable monitoring of the individual passage of
the oil and water phases via external scintillation detectors
mounted on the inlet and water/oil exit lines.
Since tracers were injected into the inlet stream as an
instantaneous pulse (Dirac delta function), the RTD of the
vessel can be obtained from the outlet concentration distribution directly (Levenspiel, 1999). At the vessel outlet, the
outlet concentration at time t, c(t) is measured, where t 0
at the instance of tracer injection at the inlet. The exit age distribution or residence time density function, E(t), is defined
as the fraction of elements leaving with ages between t and
t dt (Danckwerts, 1953). If the tracer is an instantaneous
pulse, E(t) can be obtained by normalizing the exit concentration with respect to the area under the c(t) curve thus:
E(t) 1
0

c(t)
c(t) dt

(1)

Hence
1

E(t) 1

(2)

and it follows that


1

c(t) dt

m
Q

The upper limit in the integrals in the above equations can be


replaced by some finite time, T, beyond which no more tracer
can be detected. It hence follows from Danckwerts (1953)
that the MRT can also be found from
MRTA

where m is the mass of tracer injected and Q is the volumetric


flow rate of the organic or the aqueous phase. Since E(t) is
normalised, it is not necessary to know m, although if it is
known, the mass balance can be checked. The mean residence time (MRT), is given by (Levenspiel, 1999).
1

tE(t) dt

(4)

c(t)
c(t) dt

(5)

Using

we obtain
1
MRT 01
0

tc(t) dt
c(t) dt

(7)

where VA is the active vessel volume through which each


respective phase flows.
Experimental Parameters
The RTD measurements were made for both oil and aqueous phases over a range of flow rates for the separators
described above. Operating parameters, vessel geometries
and MRTs calculated from both equations (6) and (7) are
given in Table 1, together with heights of the gas oil and
oil water interfaces obtained from the interface level indicators installed on each vessel. The MRTs for either the
organic or aqueous phase obtained from equation (7)
have been calculated using the measured volumetric flow
rate and the volume of the vessel occupied by the respective phase. The volume of each phase was determined by
calculating the cross-sectional area occupied by each
phase on the basis of the measured interface heights (illustrated in Figure 2) and then multiplying by the vessel length
to obtain a volume.
Table 1 shows that the values of MRT calculated via
both methods are different. This indicates some form of
error in the experiments or model assumptions since, by
continuity, the values should be the same. This can be
attributed either to errors in the interface height measurements or blockage of the vessel active volume due to deposition of solids or internals. These effects are illustrated by
a fractional dimensionless volume, VD, in Table 1, which
represents the actual volume occupied by each phase
based on the measured MRT, V, divided by the volume
each phase is assumed to occupy based on interface
height measurements, VA, that is,
VD

E(t) 1

VA
Q

(3)

MRT

1385

(6)

MRT
V=Q
V

MRTA VA =Q VA

(8)

MATHEMATICAL MODEL
Alternative Path Model (APM)
The APM is based on splitting the flow behaviour in the
vessels into a series of zones and includes two alternative
paths for the fluid to travel that have different time constants. Possible velocity profiles in the vessel and flow
zones are shown in Figure 3a. In the formulation of the
APM, the presence of the rag layer (the layer between
the two liquid phases containing one phase dispersed in
the other) is neglected. When systems are described in
this way, the model can be derived mathematically by
using transfer functions from Laplace Domain descriptions (Levenspiel, 1999; Luyben, 1990). A block diagram
of the APM is given in Figure 3b.
The APM has six adjustable parameters for each phase:
. time constant in inlet mixing (CSTR) zone, t1;
. time constant of a stirred tank in each series, t2, t3;

Trans IChemE, Part A, Chemical Engineering Research and Design, 2004, 82(A10): 13831390

1386

SIMMONS et al.
Table 1. Mean residence time and APM data for the field separators

MILNE POINT

Vessel diameter (m)


Vessel L/d (2)
Gasoil int. ht (m)
Oilwater int. ht (m)
Sand depth (m)

3.6
6.8
1.575
0.914
0.40

1.575
0.939
0.40

1.473
0.863
0.40

1.473
0.863
0.40

1.575
0.977
0.40

1.575
0.813
0.40

ORGANIC PHASE
Volumetric flow (m3/s)
Mean velocity (m/s)
MRT (s) from RTD equation (6)
MRTA (s) from equation (7)
Dimensionless volume, VD (2)
Sec. peak no., F (2)

0.073
0.032
680
753
0.90
1.35

0.076
0.035
621
698
0.89
1.56

0.074
0.036
668
675
0.99
0.96

0.070
0.034
622
714
0.87
1.46

0.074
0.036
854
678
1.25
0.16

0.070
0.027
730
894
0.82
0.51

AQUEOUS PHASE
Volumetric flow (m3/s)
Mean velocity (m/s)
MRT (s) from RTD equation (6)
MRTA (s) from equation (7)
Dimensionless volume, VD (2)
Sec. peak no., F (2)
Fractional mixed volume, D (2)

0.033
0.023
1585
1050
1.51
0.48
0.30

0.033
0.022
1632
1108
1.47
0.38
0.23

0.033
0.026
1152
932
1.24
0.15
0.26

0.037
0.029
1343
831
1.61
0.26
0.27

0.037
0.023
1359
1068
1.27
0.04
0.40

0.032
0.028
1421
845
1.68
0.00
0.40

10

11

KINNEIL
Vessel diameter (m)
Vessel L/d (2)
Gasoil int. ht (m)

3.05
4.03

ORGANIC PHASE
Volumetric flow (m3/s)
Mean velocity (m/s)
MRT (s) from RTD equation (6)
MRTA (s) from equation (7)
Dimensionless volume, VD (2)
Sec. peak no., F (2)
Fractional mixed vol., D (2)

1.53

1.53

1.53

1.53

1.53

0.14
0.038
300
322
0.93
0.00
0.805

0.23
0.063
198
172
1.15
0.00
0.727

0.19
0.052
240
237
1.01
0.05
0.835

0.26
0.071
130
154
0.84
0.24
0.742

0.17
0.046
269
223
1.21
0.32
0.862

MAGNUS
12

ULA
13

14

15
2.64
2.8
1.19
0.87

Vessel diameter (m)


Vessel L/d (2)
Gasoil int. ht (m)
Oilwater int. ht (m)

3
3.3
1.15
0.675

1.15
0.695

3.3
3.0
1.90
0.87

ORGANIC PHASE
Volumetric flow (m3/s)
Mean velocity (m/s)
MRT (s) from RTD equation (6)
MRTA (s) from equation (7)
Dimensionless volume, VD (2)
Sec. peak no., F (2)

0.107
0.082
191
121
1.58
0.03

0.109
0.087
191
113
1.69
0.05

0.123
0.037
162
265
0.61
2.82

0.033
0.040
173
184
0.94
0.81

AQUEOUS PHASE
Volumetric flow (m3/s)
Mean velocity (m/s)
MRT (s) from RTD equation (6)
MRTA (s) from equation (7)
Dimensionless volume VD(2)
Sec. peak no., F (2)
Fractional mixed vol., D (2)

0.064
0.053
194
184
1.05
0.02
0.58

0.062
0.050
191
198
0.96
0.05
0.61

0.143
0.079
284
124
2.30
0.63
0.47

0.060
0.038
178
193
0.92
0.46
0.70

. flow fraction through each path, f, is defined as the flow


fraction through path corresponding to t3;
. number of stirred tanks in each path, N.
The transfer function and analytical solution of this
model are as follows:
"



N #
1
1
1
(1  f )
(9)
f
G(s)
t1 s 1
t2 s 1
t3 s 1

(1  f )t1N1 t=t1
e
 (1  f )
(t  1  t2 )N
N
X
t1Ni ti1
et=t2
N1i i1
(
t

t
)
t
(i

1)!
1
2
2
i1
N
X
f t1N1 t=t1
t1Ni ti1

e

f
N1i i1
(t1  t3 )N
t3 (i  1)!
i1 (t1 t3 )
(10)

G(t)

Trans IChemE, Part A, Chemical Engineering Research and Design, 2004, 82(A10): 13831390

RESIDENCE TIME DISTRIBUTIONS IN OIL WATER SEPARATORS

1387

since the shape of the RTD becomes rapidly insensitive


to large N.
From these model parameters, it is possible to derive parameters that can be related to the hydrodynamics in the
vessel. Following Danckwerts (1953), the time constants
obtained from the APM can be used to back-calculate the
volume of the tank occupied by the liquid. First, it is possible to calculate values of MRT for the organic and aqueous
phases respectively, tmo, tmw. where o oil phase and w
aqueous phase
tmo t1o fo t2o (1  fo )t3o 
tmw t1w fw t2w (1  fw )t3w 

Figure 2. Illustration of volume occupied by each phase.

The total volume can then be calculated by multiplying by


the individual phase flow rates.
Vtotal Qo tmo Qw tmw

The parameters in the model were fitted to the experimentally determined exit age distributions, E(t), using the Simplex method. A full description of the application of the
Simplex method to this problem is given by Komonibo
(2002). A fixed value of N 50 was used for both paths,

(11)
(12)

(13)

The total size of the inlet mixing zone, Vmix can be estimated similarly.
Vmix Qo t1o Qw t1w

Figure 3. (a) Possible flow zones within separator; (b) block diagram of Alternative Path Model (APM).

Trans IChemE, Part A, Chemical Engineering Research and Design, 2004, 82(A10): 13831390

(14)

1388

SIMMONS et al.

The Fractional Mixed Volume, D, is defined as the fractional amount of the vessel active volume occupied by turbulent mixing at the inlet, that is,
D

Vmix
Vtotal

(15)

Values of fractional mixed volume close to unity indicate


high turbulence within the vessel, which would impede
settling. Values close to zero indicate a plug-like flow. If
it is assumed that the mixing zone occupies the total
volume of the vessel up to a certain point, after which
there is a distinct water oil interface, then it can be
simply derived that
D

lmix
l

(16)

where lmix is the length of the mixing zone and l is the


active length of the vessel.
As described later, a feature noticeable on several of the
exit age distributions produced was the presence of a secondary peak. A secondary peak number, F, is proposed
that can be calculated from parameters in the APM.


t3
1
(17)
Ff
t2
For full details of the physical basis of the model and the
development of the model parameters, the reader is directed
to the paper of Simmons et al. (2002).

RESULTS AND DISCUSSION


Table 1 shows the values of MRT calculated from
equations (6) and (7) and fractional dimensionless
volume, VD. For the Kinneil and Magnus vessels, the
values of VD are close to unity, which indicates accurate
measurement of interface height and an absence of
blockages within the vessels. However, this is not the
case for the Milne Point and Ula vessels where VD deviates
both above and below unity. For the Milne Point vessel, VD
is below unity for the organic phase and above unity for the
aqueous phase. This would seem to indicate a systemic
underprediction of the height of the oil water interface,
since the measured MRT of the oil and aqueous phases
are respectively shorter and longer than expected. Interface
height measurements are known to be prone to error due to
the presence of a rag layer between the oil and water phases
and this effect, although not measured, may be attributable.
This is also the case for the Ula vessel (Expt. 14), although
here the effect is even more marked. This result raises an
important and sometimes neglected issue in the control of
such separators, since the calculation of MRT for each
phase is commonly done by plant operators on the basis
of interface height information. On the basis of this data,
this can lead to errors of up to 100% (Expt. 15), with consequences for the separation efficiency.
Examples of the exit age distributions produced for both
organic and aqueous phases for the Milne Point vessel are
given in Figure 4. The injection of tracer at the vessel inlet
corresponds to time 0 on the x-axis for all these plots

Figure 4. Comparison of exit age distributions with APM for Expt. 1,


Milne Point vessel: (a) organic phase; (b) aqueous phase.

(Figures 4 6). The shape of the exit age distribution is


similar for both phases: after a short delay, there is a
rapid rise to a sharp peak, followed by a slower decay
and a long tail. This is particularly noticeable for the aqueous phase. The exact length of the tail is generally difficult
to predict because of noise in the experimental data. However, in this case there is clearly no offset between the final
and end values of the exit age distribution indicating that
the entire tracer had exited. A shoulder (indicated on
Figure 4b), or secondary peak is observable for the aqueous
phase indicating later release of tracer possibly due to a
secondary flow or dead zone somewhere in the vessel.
The fitted APM curves show a good agreement, although
there are some small discrepancies near the tail of the distribution. A check on the fit of the APM may be made by
comparing the measured MRT of each phase with the
mean residence times obtained from the model, tmo and
tmw. The agreement between these parameters was within
10% for all experiments, indicating that the fit was adequate within experimental accuracy.
Figures 5 and 6 show examples of the exit age distributions produced from the Magnus and Ula vessels. The
shapes of the curves are qualitatively very similar to

Trans IChemE, Part A, Chemical Engineering Research and Design, 2004, 82(A10): 13831390

RESIDENCE TIME DISTRIBUTIONS IN OIL WATER SEPARATORS

1389

Figure 5. Comparison of exit age distributions with APM for Expt. 12,
Magnus vessel: (a) organic phase; (b) aqueous phase.

those observed in Figure 4 except that no secondary peaks


are observed for the Magnus vessel. There is similar good
agreement with the APM. The qualitative similarity
between the distributions produced from different vessels
highlights clearly the need for a modelling approach,
since the model parameters allow subtle differences
between the E(t) curves to be identified.
The presence of secondary peaks on the exit age distributions may be due to the presence of secondary flows.
Nilsen and Davies (1996) and Simmons et al. (2002) postulated that these flows may be caused by high differential
velocities between the phases (Figure 3a). Recirculation
zones then develop, retarding the release of one or more
of the phases. Hence, a greater amount of secondary
peaks would be expected at higher differential velocities.
For the Milne Point vessel, there is a general increase in
the secondary peak number, F, with increasing differential
velocity, as shown in Table 1 and Figure 7. Insufficient
experimental data are available to compare trends in the
other vessels. The presence of a rag layer may also cause
similar effects and this needs further investigation.
The variation of fractional mixed volume with water cut
is shown in Figure 8. Since the minimization of mixing and
turbulence would be expected to ease the separation of the
phases, this parameter can be used as a measure of the separation efficiency of the vessels. Upon examination of the
values, fractional mixed volumes for Ula and Magnus are
approximately between 0.5 and 0.8. Values for Milne
Point are approximately 0.3. This suggests that the considerably more sophisticated internals of the Milne Point
vessel are having a desirable effect in reducing the overall
turbulence in the vessel and also demonstrates that this

Figure 6. Comparison of exit age distributions with APM for Expt. 14, Ula
vessel: (a) organic phase; (b) aqueous phase.

parameter can be used to characterize the relative levels


of turbulent mixing. Another important factor is the
vessel L/d ratio. The Milne Point vessel has a high L/d
of nearly 7, hence the flow has more length over which
the flow can stabilize compared with the other vessels,
which have L/d ratios from 2.8 to 4. This longer design
is advantageous. The values of fractional mixed volume
(Table 1) obtained for the Kinneil Separator are very
high, approximately 0.7 0.9. This shows that there is still

Figure 7. Variation of secondary peak number, F, with differential phase


velocity for Milne Point vessel.

Trans IChemE, Part A, Chemical Engineering Research and Design, 2004, 82(A10): 13831390

1390

SIMMONS et al.
Greek symbols
t
time constant, s
Subscripts
o
w

organic phase
aqueous phase

REFERENCES

Figure 8. Variation of fractional mixed volume, D, with water cut.

considerable mixing in the liquid phase despite internal baffling. This could be due to the rapid rise of bubbles of gas in
the oil, driven by buoyancy effects arising from large density differences between the gas and oil phases since this
vessel is used primarily for gas oil separation.

CONCLUSIONS
The alternative path model (APM) has been applied to
obtain parameters describing the hydrodynamics and separation performance of field separators from residence time
distribution (RTD) data. The model was fitted to the experimental exit age distributions using the Simplex method
and good agreement was obtained. For one of the vessels
used, the secondary peak number, F (an indication of the
presence of dead zones and re-circulatory effects) was
found to generally increase with increasing differential velocity between the organic and aqueous phases. Values of
fractional mixed volume, D, are lowest for vessels containing baffles at the inlet and down the body of the vessel to
smooth the flow and lower turbulence. A long vessel, that
is, high L/d, is advantageous. The effect of the baffles is
reduced for gas oil separation, perhaps due to the presence
of rapidly rising disengaging gas bubbles, which act to
increase the local turbulence.
NOMENCLATURE
c
d
D
E
f
F
L
lmix
m
MRT
MRTA
N
Q
t
V
VA
VD

concentration, kg m23
vessel diameter, m
fractional mixed volume
exit age distribution
flow fraction through each path
secondary peak number
vessel length, m
length of mixing zone, m
mass of injected tracer, kg
measured mean residence time, s
mean residence time (volume/volume flowrate), s
number of stirred tanks in each path
volume flowrate, m3 s21
time, s
phase volume, m3
phase volume based on interface height measurements, m3
dimensionless volume

Bailes, P.J. and Larkai, S.K.L., 1981, An experimental investigation into


the use of high voltage D.C. fields for liquid phase separation, Trans
IChemE, Part A, Chem Eng Res Des, 59: 229237.
Bailes, P.J. and Larkai, S.K.L., 1982, Liquid phase separation in pulsed
DC fields, Trans IChemE, Part A, Chem Eng Res Des, 60: 115121.
Danckwerts, P.V., 1953, Continuous flow systemsdistribution of residence times. Chem Eng Sci, 2: 113.
Dick, D., 1997, Private communications: Ula separators trials on HP and
test separators 1995, BP internal report no. XFE/014/95. Milne Point
unit slug-catcher/separator tests 1997, BP internal report no. SPR/
TRT/020/97. Performance trials on F801 separator, BP Kinneil,
Grangemouth 1997, BP internal report no. 10/EC97/149.
Gerunda, A., 1981, How to size vapour liquid separators, Chem Engng,
88(9): 8184.
Hansen, E.W.M., Heitmann, H., Laksa, B., Ellingsen, A., Ostby, O.,
Morrow, T.B. and Dodge, F.T., 1991, Fluid flow modelling of gravity
separators, in Proceedings of 5th International Conference on Multiphase Production, 1991 Cannes, Burns, A.P. (ed), pp. 364381.
Ingersoll, A.C., 1951, Fundamentals and performance of gravity
separation a literature review, in Proceedings, API 31st Meeting, Division of Refining, Vol. 31M(III).
Komonibo, E., 2002, Interpretation of the flow characteristics of a primary
oilwater separator from the residence time distribution, MRes Thesis,
University of Nottingham.
Levenspiel, O., 1999, Chemical reaction engineering, 3rd edition, (John
Wiley and Sons, USA).
Luyben, W.L., 1990, Process modelling, simulation and control for chemical engineers, 2nd Edition (McGraw-Hill International Publications,
USA).
Meon, W., Rommel, W. and Blass, E., 1993, Plate separators for dispersed
liquid/liquid systems: hydrodynamic coalescence model, Chem Eng Sci,
48(1): 159168.
Mohamad Nor, M.I. and Wilkinson D., 1998, Experimental and CFD
studies to improve oilwater separation in horizontal separators, in Proceedings of 1998 IChemE Research Event, Newcastle, England,
CDROM.
Nilsen, F.P. and Davies, G.A., 1996, What is wrong with primary
separators? (Christian Michelsen Research, Bergen, Norway; Department of Chemical Eng, UMIST, Manchester, UK).
Rashad, M.A., Davies G.A., Bos A., 1997, CFD modelling of flow
distribution methods in primary separators, J Comput Fluid Dynamics,
10(4): 374390.
Rowley, M.E. and Davies, G.A., 1988, The design of plate separators for
the separation of oil/water dispersions, Trans IChemE, Part A, Chem
Eng Res Des, 66(4): 313322.
Simmons, M.J.H., Wilson, J.A. and Azzopardi B.J., 2002, Interpretation of
the flow characteristics of a primaryoilwater separator from the residence time distribution, Trans IChemE, Part A, Chem Eng Res Des,
80(A): 471481.
Wilkinson, D., and Waldie, B., 1994, CFD and experimental studies of
fluid and particle flow in a horizontal primary separator, Trans
IChemE, Part A, Chem Eng Res Des, 72(A): 189196.
Wilkinson, D., Waldie, M.I., Nor, M., and Lee, H.Y., 2000, Baffle plate
configurations to enhance separation in horizontal primary separators.
Chem Eng J, 77: 221226.

ACKNOWLEDGEMENTS
M.J.H. Simmons would like to acknowledge funding by the School
of Chemical, Environmental and Mining Engineering, University of
Nottingham, and BP Exploration
The manuscript was received 10 November 2003, and accepted for
publication after revision 14 July 2004.

Trans IChemE, Part A, Chemical Engineering Research and Design, 2004, 82(A10): 13831390

Вам также может понравиться