Sei sulla pagina 1di 9

OTC 17386

Emulsion Rheology - Theory vs. Field Observation


R.C.G. Oliveira and M.A.L. Gonalves, Petrobras

Copyright 2005, Offshore Technology Conference


This paper was prepared for presentation at the 2005 Offshore Technology Conference held in
Houston, TX, U.S.A., 25 May 2005.
This paper was selected for presentation by an OTC Program Committee following review of
information contained in a proposal submitted by the author(s). Contents of the paper, as
presented, have not been reviewed by the Offshore Technology Conference and are subject to
correction by the author(s). The material, as presented, does not necessarily reflect any
position of the Offshore Technology Conference, its officers, or members. Papers presented at
OTC are subject to publication review by Sponsor Society Committees of the Offshore
Technology Conference. Electronic reproduction, distribution, or storage of any part of this
paper for commercial purposes without the written consent of the Offshore Technology
Conference is prohibited. Permission to reproduce in print is restricted to a proposal of not
more than 300 words; illustrations may not be copied. The proposal must contain conspicuous
acknowledgment of where and by whom the paper was presented. Write Librarian, OTC, P.O.
Box 833836, Richardson, TX 75083-3836, U.S.A., fax 01-972-952-9435.

Abstract
Petroleum companies are familiar with the problems caused by
the formation of water-in-crude oil emulsions, especially with
those related to the increment of the crude oil apparent
viscosity and their impact on the oil production. Many studies
have been conducted to establish the existing relationship
between the crude oil emulsions apparent viscosity and their
water content. However, there is a lack of information in the
literature concerning the actual impact of the emulsions on the
pressure drop trough pipelines and on its consequence, the
reduction of crude oil production.
The thermo-hydraulic calculation of a producing oil stream
is usually based on black-oil models for PVT properties and
on multiphase flow correlations or models suitable for
pressure drop prediction. These models and correlations
demand the knowledge, amongst other parameters, of the
viscosity of the liquid phase as an input data. Usually, the
values of this property are measured with rheometers using
standardized procedures. However the thermo-hydraulic
calculation becomes more complex in the case of water-in-oil
emulsions. The question is how to get a correct viscosity to
use for the pressure drop calculation and thus, how to prepare
a synthetic fluid that represents the actual emulsion. The
objective of this study was to understand the state of the art
relative to the emulsion formation and rheology. The main
rheological models that are applicable to emulsions are
presented, as well as the methodologies for the determination
of its parameters in the laboratory. This paper also investigates
the usual correlations applied for the flow of gas and water-inoil emulsions through pipelines. Actual data from field were
used to allow the evaluation of the results obtained through the
simulations using the proposed procedure. This information is
relevant to establish how the presence of water-in-crude oil
emulsions can interfere with the technical and economical
viability studies.

Introduction
Water-in-oil (W/O) emulsions are very common in the
petroleum industry in particular at the upstream operations.
They form naturally during the crude oil production and the
water content can be as high as 60% by volume.
The presence of W/O emulsions may have a strong impact
on the crude oil production, especially in offshore conditions.
In such kind of systems the temperature of the crude oil varies
widely along the flow from the reservoir to the platform
storage tanks. For example in Campos basin, where most of
the Brazilian crude oil is produced, typically the temperature
of the oil gradually decreases from 80 C at the bottom of the
well-bore, located 3,000 m below the seabed, to about 60 C at
the top of the well-bore, located 1,000 m below the sea level,
where the sea water temperature may vary from 4 C to 10 C.
In order to reach the storage tank, the crude oil has to flow for
several hundreds meters through a pipeline in a cold
environment. The contact with the cold seawater imposes a
major decrease in the crude oil temperature; hence crude oil
arrival temperatures below 30 C have been frequently
reported at Campos Basin.
In general, those water-in-oil emulsions are stable and
form spontaneously due to the presence of natural surfactants
existing in the crude oil phase. It is known that the viscosity of
a W/O emulsion is strongly augmented by increasing its water
volume fraction and by decreasing the temperature. It is also
dependent on the water droplet size distribution and therefore
on the shear stress that acts thru the production process.
Correlations can be found in the literature [1, 2] between the
relative viscosity (r) of the water-in-crude oil emulsions and
their water volume content and oil phase density, most of them
based on the well-known ASTM equation [3]. Nevertheless,
there is a lack of experimental data available in the literature
as far as the effect of the droplet size distribution on the crude
oil viscosity is concerned. The aim of the present study was to
determine how the water-in-crude oil emulsions rheology is
affect by the shearing conditions imposed during the
emulsions generation step. The main rheological models that
are applicable to emulsions are presented, as well as the
methodologies for the determination of its parameters in the
laboratory. This paper also investigates the usual correlations
applied for the flow of gas and water-in-oil emulsions through
pipelines. Actual data from field were used to allow the
evaluation of the results obtained through the simulations
using the proposed procedure. This information is relevant to
establish the W/O emulsions influence on the well
productivity.

OTC 17386

Models for the emulsions viscosity prediction


The viscosity () of mono-disperse water-oil emulsions with
similar densities of phases and low interfacial tension, under
steady-state condition, is usually a function of the following
variables: temperature (T), dispersed-phase volume fraction
(), viscosity of the continuous phase (C), viscosity of the
disperse phase (D), process shear rate (Df), droplet radius (R),
and pressure (P). At constant pressure and shear rate, the main
variables that can affect the viscosity of a water-in-crude oil
emulsion are the temperature and the water volume fraction.
However, the droplet size distribution also becomes an
important variable for concentrated emulsions.
Models at constant temperature. Most of the existing
models are based on the parameter called relative viscosity
(r), defined as the ratio between the emulsion and the oil
continuous phase viscosities:
r = /C .......................................................................................................... (1)
Many years ago, Einstein developed a thermodynamic
model for colloids and proposed that the relative viscosity of
diluted emulsions increases linearly with the volume fraction
of dispersed phase:

r = 1 + 2.5 ............................................................. (2)


Later, Taylor studying the emulsions based on the Fluid
Mechanics principles proposed another expression, also valid
for diluted solutions and spherical drops. This relation is
similar to that of Einstein except that it introduces the
influence of the viscosities of both phases:

r = 1 + [2.5 (k+0.4)/((k+1)] ................................... (3)


where k is the ratio between the dispersed and the continuous
phase viscosities:
k = D / C................................................................. (4)
Taylors model does not consider the deformation of the
drops and the interactions between them, which can be
significant especially for concentrated emulsions. To take into
account these effects, Choi and Schowalter [4] and Yaron and
Gal-Or [5] have proposed correction factors (f) as a function of
the dispersed-phase volume fraction:

r = 1 + f(1/3) ......................................................... (5)


Woelflin [6] investigated experimentally the viscosity of
emulsions and classified them according to three groups
named loose, medium and tight. Although this author
did not present a model to describe the behavior of these
emulsions, the experimental data has been used in several
computational codes by means of numerical data fit. In this
work Woelflin data is fitted through:
r = exp(a 2 + b ) ................................................... (6)
where a and b are parameter dependent on the type of
emulsion (loose, medium or tight).
Phan-Thien and Pham [7] developed another model for
concentrated emulsions using the approach of the effective
medium average. Starting from the equation of Taylor they
developed a differential equation for the viscosity of

concentrated emulsions at low capillary numbers whose


solution for the boundary conditions ( = 0 = c) is:

r2/5 [(2r+ 5 k)/(2+5k)]3/5= (1-)-1 ............................(7)


Pal [8, 9] has recently reviewed and evaluated several
theoretical viscosity models for dilute and concentrated
emulsions. All the reviewed models predict the relative
viscosity of emulsions as a function of dispersed-phase
volume fraction and the ratio between the viscosity of
dispersed-phase and that of the continuous phase.
Unfortunately, these models do not consider the strong
effect of the temperature on the W/O emulsions viscosity.
They just describe the variation of viscosity as a function of
the dispersed-phase volume fraction, or the ratio between the
viscosity of dispersed-phase and that of the continuous phase.
The temperature effect. Ronningsen [2] has proposed the
following correlation for the viscosity of W/O emulsions as a
function of the dispersed-phase volume fraction and
temperature:
ln e = a1+ a2T + a3 + a4T ....................................(8)
where a1, a2, a3 and a4 are constants that are function of the
process shear rate.
This correlation is based on the exponential relationship
between the viscosity and the dispersed-phase volume fraction
and was obtained through the analysis of experimental data at
different temperatures and shear-rates. For crude oils and
theirs fractions, ASTM [3] recommends an equation for the
variation of the kinematic viscosity with the temperature:
ln ln Z = A - B ln T ...................................................(9)
where A and B are constants for a specific emulsion, T is the
absolute temperature and Z is done by equation 9.
Z = + 0.7 for larger than 2.0 cSt .......................(10)
The ASTM method is widely accepted in the petroleum
industry; however its use is limited only to crude oil and their
fractions until today.
Recently Farah et al. [10] have proposed a model based on
the equation 9. This model assumes that the values of the
constants A e B of the ASTM equation exhibit a linear
relationship with the dispersed-phase volume fraction.
A = k1 + k2 () ............................................................(11)
B = k3 + k4 () ...........................................................(12)
Substituting equations 11 and 12 into 9:
ln ln( + 0,7) = k1 + k2 () + k3 ln T + k4 () ln T ....(13)
For a fixed shear rate, equation 13 permits the prediction of
the water-in-crude oil emulsions viscosity at any temperature
and at different water volume fractions.
Emulsion creation and stability
In order to allow the prediction of the water-in-crude oil
emulsion viscosity before its appearance in a production
system, it is necessary to create synthetic emulsions in lab
conditions. In this work, the synthetic W/O emulsions are

OTC 17386

prepared in such a way to mock, in terms of droplet size


distribution and mean diameter, those obtained from samples
taken from the first stage separator in a similar production
field.
A homogenizer is used for the preparation of the W/O
emulsions with 10%, 20%, 30%, 40% and 50% by volume of
salty water. The emulsions are generated using different speed
levels, from 6,000 rpm (low shear, LS; 15,500 s-1) up to
10,000 rpm (high shear, HS; 25,800 s-1), during three minutes.
The 8,000 rpm speed, considered an intermediate shear level
(IS; 20,600 s-1), is the one that, based on our experience,
permit to reproduce the formation of W/O emulsions similar to
those collected at Petrobras offshore production systems. As
soon as the emulsions are prepared they are submitted to a
standard stability test. This test consists on leaving the
emulsions to rest by 4 hours at 60 C. The W/O emulsions
considered unstable in this test are discarded.
The viscosities of the dehydrated crude oil and of their
W/O emulsions are determined using a rheometer HAAKE,
model Rheostress-1, coupled to a microcomputer. The
rheological data are collected at different water volume
fractions; shear rates and temperatures, above and below the
crude oil wax appearance temperature (WAT).
Lab observations
The effect of the shear condition and the water volume
fraction on the W/O emulsions droplet size and apparent
viscosity. Figure 1 shows the effect of the shear condition on
the W/O emulsions droplet size distribution. As can be seen
from this figure, the shear condition applied during the
emulsion generation plays an important role on the emulsions
droplet size distribution. It is observed that as the shear energy
increases, the size of the internal phase droplets tends to
decrease. This effect causes an influence on the emulsions
apparent viscosity, leading to the establishment of a different
rheological curve as shown in Figure 2. At constant water
volume fraction, this behavior is related to the increment of
the total surface area of the droplets dispersed in the oil phase.
The effect of the water volume fraction on the W/O
emulsions droplet size distribution and mean diameter are
shown in Figure 3 and 4, respectively. The results show that
increasing the water volume fraction provokes the increment
of the emulsions droplets mean diameter. This behavior is
related to the droplets close packing. As the droplets approach
each other they tend to coalesce and to form a packing in
which small droplets occupy the free space between large
droplets. Nevertheless, Young-Laplace equation:

P12 = P1 P2 = 2. .(

1
1

) ................. (14)
R1 R2

where, P12 is the difference of internal pressure from droplets


of different size and is the interfacial tension, shows that due
to the interface curvature the pressure inside small droplets is
higher than that of large droplets. This phenomena leads to the
reduction of the emulsions stability as the water phase tends
to transfer spontaneously from the small to the large droplets.
The effect of the water volume fraction on the W/O
emulsions relative viscosity. The effect of the water volume

fraction on the W/O emulsions relative viscosity is presented


in Figure 5. It can be verified that the W/O emulsions relative
viscosity increases almost linearly up to water volume fraction
values of 0.2. Above this value the W/O emulsions relative
viscosity increases exponentially, as predicted by the PhanThien and Pham model [6].
As shown in Figure 6, the relative viscosity of the W/O
emulsions decreases very fast when the temperature is reduced
below the WAT ( 25C). The results present in Figure 7 also
show that the W/O emulsions exhibit a slightly NonNewtonian behavior at temperatures above WAT. However,
below WAT those emulsions behave as strong shear thinning
fluids, especially at high values of water volume fraction.
These results indicate that the presence of wax crystals,
dispersed on the oil phase, affect the polar interaction between
the water droplets leading to a reduction on the W/O
emulsions viscosity.
Viscosity prediction for the water-in-crude oil emulsions.
As mentioned in section 2.2, Farah et al. [10] have proposed
the equation 13 for the prediction of the water-in-crude oil
emulsions viscosity at any temperature and dispersed-phase
volume fraction. This model assumes that the values of the
constants A e B of the ASTM equation exhibit a linear
relationship with the dispersed-phase volume fraction. This
equation has proved to give reasonable correlation with actual
data as shown in Figure 8.
The effect of the gas phase. The viscosities of two samples of
actual dead oil emulsion taken from field were measured
under atmospheric pressure with temperature of 21 C and
shear rate of 50 s-1. The results are shown in Table 1.
Basically, the viscosity of the dead oil emulsions differs
from the viscosity the live oil emulsion due to the presence of
dissolved gas. An experimental procedure was carried out to
recombine oil and gas so that the viscosity of a live oil
emulsion could be measured. Table 2 presents a comparison
between the measured and estimated viscosities calculated by
different prediction models. The viscosity of the dead oil
without emulsion is also presented. A similarity is noticed
between the values measured and calculated with the three
models of Woelflin. One must, however, to consider that the
sample was taken from the field and transported to the
laboratory, and during this process it may had been submitted
to a process of oxidation and aging.
The correlation of Ronningsen using the parameters for
shear rate of 500 s-1 provided results that best fit the laboratory
measurements. Although the shear rate typically found in field
conditions range from 100 s-1 to 200 s-1, the constants of the
model for higher shear rate usually grant a better result.
The importance of the emulsion rheology on the
multiphase flow calculation
Multiphase flow calculation is usually performed for field
design purposes. During the operation of oil fields, the use of
numerical flow simulation also allows the diagnosis of
possible flow assurance problems in wells and flowlines. This
is done through the comparison between measured values and
the results of simulation, and a large difference between

OTC 17386

calculated and actual values may indicate some flow assurance


problem.
The knowledge of emulsion rheology is usually one of the
main concerns while gathering data for numerical simulations.
The first point to be highlighted is that the models and
correlations available in the literature do not consider the
Non-Newtonian behavior of the water-in-crude oil emulsions.
Moreover, there may be other sources of uncertainties related
to the input data needed for numerical simulations that must be
also addressed.
To evaluate the influence of the uncertainties of the input
parameters on the calculation of pressure drop in multiphase
lines, a study was carried out using PETROBRAS in house
multiphase flow simulator MARLIM [11]. This code features
most of published multiphase flow pressure drop correlations
coupled with thermal calculation and considers the Black Oil
model for phase behavior.
This analysis takes into account only the influence of the
uncertainties of the input parameters considered for the
simulation. The uncertainty of the correlation used for
pressure drop prediction was not considered here. In fact, this
uncertainty can be much higher than the obtained value.
However, to quantify this uncertainty is a hard task and the
difficulty starts in the original publications of those flow
correlations that almost never present this information.
Anyway, the engineer usually has to choose the correlation
that best fits a given situation, and does not have much to do
regarding this task.
A discrete model for tubing and flowlines representing an
actual oil well was considered with all boundary conditions
applied. The Hagedorn-Brown [12] and Beggs & Brill [13]
correlations were used for vertical and inclined flow
computation, respectively. The Bottom Hole Flowing Pressure
(BHFP) was chosen as a parameter for comparison.
The total uncertainty U due to input data was estimated
through the equation:
U =k

U
ci i
ki
i

................................................... (15)

where the term inside the parenthesis expresses the


contribution of each input parameter to the uncertainty.
The calculated BHFP value for a given particular case of a
well, located at Campos Basin Brazil, was equal to 230.2
kgf/cm2. Using estimated input data uncertainties and
Equation 15, the overall uncertainty obtained was equal to
5.6 kgf/cm2, or 2.4%.
The rate between each parcel and the total square sum
gives a good representation of the contribution of each
parameter to the overall uncertainty. This information allows
the determination of the parameters that imposes higher
uncertainty on the calculation of the BHFP.
The input data which uncertainties are more relevant for
the calculated BHFP value are the formations gas-oil ratio
(responsible for 24% of the total BHFP uncertainty), the water
cut (54%), the liquid phase viscosity (5%) and the pipeline
roughness (7%).

This means that, for the purpose of calculating multiphase


flow through computer simulators, the question resumes into
how much water is actually being produced and how this
amount of water is partitioned as free and emulsioned water.
The influence of different models for emulsion
viscosity calculation
The liquid phase (emulsion) viscosity is among a relevant
parameter for the computational flow simulation. Therefore,
the selection of the proper model of emulsion viscosity is
important and must be investigated.
Table 3 presents the viscosity of W/O emulsions predicted
by the different models available at MARLIM [11] (Black Oilnot considering emulsion formation, Woelflin, Farah Model
and Ronningsen) and the corresponding values of BHFP. The
purpose here is to evaluate the impact of the adopted viscosity
model, therefore the same multiphase flow pressure drop
correlations were used for all cases (Hagedorn & Brown [12]
for vertical and Beggs & Brill [13] for horizontal pipes). Also
presented in Table 3 are simulations results obtained with
OLGA [14], conjugated to the PVTSim (calculation of the
fluids properties).
These BHFP data for different models were calculated for
two Campos Basin wells, that have a Pressure Downhole
Gauge installed. The calculated values can be compared to the
measured ones that are 243 kgf/cm2 for Well A and 230
kgf/cm2 for Well B. It is important to mention that these
values are subject to measurement uncertainties themselves.
The analysis of this data indicates that even though there is
a significant discrepancy between the emulsion viscosity
values depending on the selected correlation, the results of
BHFP only changes slightly. The values calculated by all
models overpredict the value of the BHFP, except those that
do not consider the emulsion formation. Nevertheless, the used
multiphase flow correlations tend to overpredict the pressure
drop, and apparently the assumption of no emulsion formation
somewhat compensates this fact.
The mechanistic model used by OLGA [14] also gave a
good forecast of the BHFP.
Although the small difference found among the calculated
and measured values, it is important to note that offshore wells
usually have very high Productivity Indices (PI), and a tiny
difference in the pressure drop could result into a large
absolute change in the production flowrates.
Pressure drop correlations for gas-emulsion flow
In order to establish the best pressure drop correlation to be
used for gas-emulsion flow, the correlations suitable for
horizontal, vertical and strongly inclined pipes were analyzed
separately. The basic difference between the flow in slightly
inclined and vertical pipelines is that in the last one the
gravitational term is dominant, while that in the horizontal and
with small inclination lines the pressure drop depends
essentially on the viscous dissipation.
A typical Campos Basin production arrangement is shown
in Figure 9. The results point that, regardless the chosen flow
correlation, most of the pressure drop (90%) occurs in the
vertical or very inclined section of the production system.
Considering this fact, the choice of the correlation to be

OTC 17386

applied to the well bore and the riser is more critical than that
applied to the horizontal flow section.
The fact that most of the pressure drop is due to vertical
sections explains the large importance found for the
formations Gas-Oil Ratio (GOR) and for the water cut in the
uncertainty analysis carried on earlier in this paper, since that
gravitational forces are by far the most relevant parameter for
vertical flow.
Analyzing the literature information one can notice that
there are quite a few published correlations suitable for gasemulsion flow in large diameter pipelines. In fact, the different
correlations provide a wide range of results that is broader
than that provided by adopting different models for the
emulsion viscosity. From a literature survey it was found that
only the Orkiszewsky [15] correlation comprises a similar
range of parameters to the field under analysis but
unfortunately this correlation was developed only for vertical
flow and does not apply to inclined pipes. When comparing
these different correlations it can be noted that they usually
tend to overpredict the pressure drop of viscous emulsions.
Pressure drop behavior related to emulsion water
cut
As the amount of produced water rises, one can expect a
corresponding increase of the emulsion viscosity, and as a
consequence the pressure drop through the producing system
should increase as well. The graphs presented in Figure 10,
obtained in three wells of the Campos Basin, however, show a
different behavior from the expected one.
In the case of Well X, for example, there is a slight
decrease in the pressure drop from zero to 30% water cut.
Above this value of water cut the pressure drop starts to
increase as expected. For the Well Y, one evidences that there
is a significant reduction of the pressure drop as the water cut
rises slowly, indicating the influence of other parameters not
presented in the graph. Finally, the Well Z presents a similar
behavior of that of the Well X, but with small changes on the
pressure drop as the water cut increases.
It must be stressed that the pressure drop data itself is
subjected to measurement errors. However, they strongly
suggest that parameters other than emulsion viscosity may be
having a decisive role in the flow.
According to the theory of single-phase flow, for high
values of Reynolds numbers - turbulent pattern - the increase
of viscosity does not impact significantly the friction factor.
The Table 4 below presents the flow patterns for three wells
from Figure 10, considering their different sections of the sub
sea production system and properties along the flow path. It
was observed the predominance of the turbulent flow, with
only one exception that presents laminar and transition
patterns. In part, this fact explains the low impact of the
emulsion viscosity on the pressure drop of most production
systems.
Conclusions
This work presented a comprehensive study involving
experimental and analytical-numerical parts, aiming to
determine the influence of the occurrence of W/O emulsions
on the production stream.

The analysis of the experimental data, obtained for


different water-in-crude oil emulsions accomplished at
different temperatures and dispersed-phase volume fractions,
allowed us to reach the following conclusions:
below the WAT, the W/O emulsions behave as shear
thinning fluids, especially at high values of water
volume fraction;
the W/O emulsions apparent viscosity and its shear
thinning behavior increase as the droplet mean
diameter diminishes and the droplet size distribution
is reduced;
a broader droplet size distribution and a larger mean
diameter is obtained with the increase of the W/O
emulsions water cut;
when the temperature of the W/O emulsions is
reduced below the WAT, their relative viscosity
decreases very fast;
the relative viscosity of the W/O emulsions increases
almost linearly up to 20% water cut. Above this value
their relative viscosity increases exponentially as
predicted by the Phan-Thien and Pham model [6];
the predicted viscosities of W/O emulsions calculated
from Farah et al. [9] and Ronningsen [2] showed a
good agreement with those obtained at lab conditions.
This paper also presents a study on the influence of the
emulsion viscosity on the multiphase flow of the produced
stream, concluding that:
the knowledge of the emulsion rheology uniquely is
not enough for pressure drop prediction. A suitable
multiphase flow model or correlation is required;
there are no published multiphase flow pressure drop
model or correlation specific for gas-emulsion flow;
the multiphase flow pressure drop tends to be over
predicted by most of published correlations;
due to the turbulent nature of the analysed wells
(which reproduce typical offshore scenarios) and the
predominance of the vertical parcel on the pressure
drop, the exact knowledge of emulsion viscosity may
have limited importance for the prediction of the
multiphase flow pressure drop;
field data correlating pressure drop with water cut
seem to suggest that other parameters beside the
viscosity may be strongly effecting the multiphase
flow behavior.
Acknowledgments
We thank Petrleo Brasileiro S/A (PETROBRAS) for the
support of this research and for the permission to publish this
paper. Thanks also to Roberto da Fonseca Junior and Dr.
Marcia C. Khalil de Oliveira for their valuable help in
computational and laboratory tasks, to Professor Luis
Fernando Alzuguir Azevedo (PUC-RJ) for the help on the
sensitivity analysis and to Dr. Geraldo Spinelli Ribeiro for the
careful review.

OTC 17386

References
1.

2.
3.
4.
5.
6.
7.
8.
9.
10.

11.
12.

13.
14
15.

Oliveira, R. C. G. , Oliveira, M. C. K. Levantamento de


correlaes para predio da reologia de emulses de diferentes
tipos de petrleos, Petrobrs Internal Report, Rio de Janeiro,
2001.
Ronningsen, H. P. Correlations for predicting viscosity of
W/O-emulsions based on North Sea Crude oils .Proc. SPE Int.
Symp. Oil Field Chem., Houston, Texas, 1995.
Annual Book of ASTM Standards Section 5, Volume 4, ASTM,
Philadelphia, PA. 2001.
Choi, S. J , Schowater. W. R. Rheological properties of nondilute suspensions of deformable particles, Physics of Fluids,
18, pp. 420 427, 1975.
Yaron, I, Gal-Or, B. Viscous flow and effective viscosity of
concentrated suspensions and emulsions, Rheologica Acta 11,
pp. 241 252, 1972.
Woelflin, W., The Viscosity of Crude-Oil Emulsions, Drilling
and ProductionPractice, API (148-53), 1942.
Phan-Thien, N., Pham, D. C. Differential multiphase models
for poly-dispersed suspensions and particulate solids, J. NonNewtonian Fluid Mech., 72, pp. 305 318, 1997.
Pal, R. A novel method to correlate emulsion viscosity data,
Colloids and Surfaces, A: Physicochemical and Engineering
Aspects, 137, pp. 275 286, 1998.
Pal R. Evaluation of theorical viscosity models for
concentrated emulsions at low capillary numbers, Chemical
Engineering Journal, 81, pp. 15 21, 2001.
Farah, M., Oliveira, R. C. G., Oliveira, M. C. K., Rajagopal, K.
A New Correlation for Estimating Viscosity of Water in Oil
Emulsions, Ibero American Conference on Phase Equilibria
and Fluid Properties for Process Design, Foz do Iguau, PR,
2002.
Almeida,A.R., Slobodcicov,I Continuous gas lift performance
analysis, World Oil, v. 219, n. 9, p. 91-96, Sept. 1998
Hagedorn, A.R. e Brown, K.E., Experimental Study of
Pressure Gradients Occurring During Continuous Two-Phase
Flow in Small-Diameter Vertical Conduits, JPT, (475-484),
April 1965.
Beggs, H.D. e Brill, J.P., A Study of Two-Phase Flow in
Inclined Pipes, JPT (607-617). May, 1973.
Bendiksen, K.H, Malnes, D., Moe, R. and Nuland S. The
Dynamic Two-Fluid Model OLGA: Theory and Application,
SPE Production Engineering, May 1991.
Orkiszewski, J., Predicting Two-Phase Pressure Drops in
Vertical Pipe, JPT, (829-838). June 1967.

OTC 17386

6.0

60
Low Shear
Intermediate Shear
High Shear
= 0.3

40

% in Volume

5.5
Droplet Mean Diameter (m)

50

30
20

5.0
4.5
4.0
3.5
3.0
2.5

10

2.0

0.1

0.2

0.3

0.4

0.5

0.6

9
Water Volume Fraction

Drop Diameter (m)

Figure 1: The effect of the shear condition on the emulsions


droplet size distribution for a typical Brazilian heavy crude oil
emulsion prepared at 0.3 water volume fraction.

Figure 4: The effect of the water volume fraction on the


emulsions droplet mean diameter for a typical Brazilian heavy
crude oil emulsion prepared at intermediate shear.
9.0
8.0

45

35

Low Shear
High Shear

30

= 0.3

Relative Viscosity

7.0

40

Apparent Viscosity (Pa.s)

T = 50C

25

6.0
5.0
4.0
3.0

20

2.0

15

1.0

10

0.0
0.0

0.1

0.2

0.3

0.4

0.5

0.6

Water Volume Fraction

0
0

10

15

20

25

30

35

40

45

50

55

Figure 5: The effect of the water volume fraction on the W/O


emulsions relative viscosity for a temperature equal to 50C and
-1
shear rate of 50 s .

Temperature (C)

Figure 2: The effect of the shear condition applied during the


emulsions generation on the apparent viscosity of a typical
Brazilian heavy crude oil emulsion.

8.5
8.0

60

Relative Viscosity

Intermediate Shear

% in Volume

50
Water
Volume
Fraction

40

0.3
0.4
0.5

30

7.5
WAT
7.0
6.5
= 0.5

20

6.0

10

5.5
0

10 15 20 25 30 35 40 45 50 55
Temperature (C)

0
0

10

11

Drop Diameter (m)

Figure 3: The effect of the water volume fraction on the


emulsions droplet size distribution for a typical Brazilian heavy
crude oil emulsion prepared at intermediate shear.

Figure 6: The effect of the temperature on the W/O emulsions


relative viscosity for a water volume fraction equal to 0.5 and
-1
shear rate of 50 s .

OTC 17386

1.05

60

Well X

50

Water Cut (%)

Power Law Index, n

1.00
0.95
0.90

30
20
10

Water
Volume
Fraction

0.85

40

0
180

190

200

0.2
0.5

0.80

220

230

Well Y

5 10 15 20 25 30 35 40 45 50 55 60 65
Temperature (C)

Water Cut (%)

50

0.70
0

240

60

WAT

0.75

210

Pressure Drop (kgf/cm)

Figure 7: The effect of the temperature and the water volume


fraction on the W/O emulsions rheological behavior.

40
30
20
10
0
180

190

200

210

220

230

240

Pressure Drop (kgf/cm )

10000

Well Z

50

Water Cut (%)

Estimated Kinematic Viscosity (cm/s)

60

R = 0,9913
40
30
20

10

1000
0
180

190

200

210

220

230

240

Pressure Drop (kgf/cm )

Figure 10: Pressure drop of three sub sea wells related to their
water cut evolution.

100
100

1000

10000

Table 1: Data from the dead oil emulsions

Actual Kinematic Viscosity (cm/s)

Emulsion Sample
Density (API)
BS&W (% v/v)
Viscosity (mPa.s)

Figure 8: Correlation involving the actual and the estimated


kinematic viscosity calculated for water-in-heavy crude oil
emulsions prepared with different water contents.

Well A
19
20
520

Well B
19
44
1373

Table 2: Comparison between measured and estimated


viscosities obtained from different models: Dead Oil (black oil),
Woefflin (Loose, medium and tight emulsion) and Ronningsen
-1
(500 s )

Figure 9: Pressure drop distribution in a typical production


system from Campos Basin - Brazil.

Sample

Actual
Value
(mPa.s)

Dead Oil

Loose
Emulsion

Medium
Emulsion

Tight
Emulsion

Ronningsen

Predicted Viscosity (mPa.s)

Well A

28,00

7,82

15,40

14,54

16,00

33,16

Well B

75,00

8,52

54,18

65,40

82,04

74,18

OTC 17386

Table 4: Reynolds and pattern of the subsea wells


Table 3: Emulsion viscosities calculated by different models and
their impact on the calculated pressure drop.

Well

WELL A
Model
Live Oil
Loose Emulsion
Medium Emulsion
Tight Emulsion
Farah Model
Ronningsen
OLGA / PVTSim
No emulsion
OLGA / PVTSim
emulsion

Viscosity (mPa.s)
50C
70C
21,53
3,71
44,36
6,77
42,06
6,39
46,21
6,95
28,46
4,73
37,71
5,76

BHFP
(Kgf/cm2)
243,31
253,12
252,29
253,66
246,18
249,18

32,72

9.84

234,41

68,77

16,93

240,2

WELL B
Model
Live Oil
Loose Emulsion
Medium Emulsion
Tight Emulsion
Farah Model
Ronningsen
OLGA / PVTSim
No emulsion
OLGA / PVTSim
emulsion

Viscosity (mPa.s)
50C
70C
16,46
5,26
104,30
33,26
124,55
39,54
207,51
50,86
87,93
16,00
87,82
27,34
20,48
194,62

6,05
55,45

BHFP
(Kgf/cm2)
235,83
254,19
255,80
257,52
252,29
252,48

Postion

Reynolds

Pattern

Bottom hole

4.761

Transition

Well head

3.756

Transition

Pipe end

4.680

Transition

Pipe start

2.241

Laminar

Riser base

2.142

Laminar

Riser top

1.880

Laminar

Bottom hole

65.677

Turbulent

Well head

46.047

Turbulent

Pipe end

47.030

Turbulent

Pipe start

15.932

Turbulent

Riser base

17.834

Turbulent

Riser top

15.188

Turbulent

Bottom hole

38.522

Turbulent

Well head

30.776

Turbulent

Pipe end

18.053

Turbulent

Pipe start

10.187

Turbulent

Riser base

11.415

Turbulent

Riser top

9.615

Turbulent

189,72
237,96

Potrebbero piacerti anche