Você está na página 1de 16

SPE 121669

Development of a Methodology for the Optimization of Dehydration of


Extra-Heavy Oil Emulsions
C. Dalmazzone, SPE, and C. Noïk, SPE, IFP, and P. Glénat and F. Dang, Total

Copyright 2009, Society of Petroleum Engineers

This paper was prepared for presentation at the 2009 SPE International Symposium on Oilfield Chemistry held in The Woodlands, Texas, USA, 20–22 April 2009.

This paper was selected for presentation by an SPE program committee following review of information contained in an abstract submitted by the author(s). Contents of the paper have not been
reviewed by the Society of Petroleum Engineers and are subject to correction by the author(s). The material does not necessarily reflect any position of the Society of Petroleum Engineers, its
officers, or members. Electronic reproduction, distribution, or storage of any part of this paper without the written consent of the Society of Petroleum Engineers is prohibited. Permission to
reproduce in print is restricted to an abstract of not more than 300 words; illustrations may not be copied. The abstract must contain conspicuous acknowledgment of SPE copyright.

Abstract
With the increasing development of heavy and extra-heavy oil fields, separation operations are becoming more and more
challenging compared to separation for conventional oil fields. For in situ bitumen Extra Heavy Oils produced thanks to
thermal process, dehydration requires solvent addition, injection of large amount of demulsifier additives, relatively high
operating temperature, and long retention times inside the separators. So in order to respect specifications on crude oil and
water quality at the lower cost, an optimization of the different parameters involved in the whole process of separation
becomes necessary.
In the case of extra-heavy oils, the presence of polar heavy components, such as asphaltenes, structured as a rigid film at
the water/oil interface, limits the coalescence phenomena and consequently limits the efficiency of separation by gravity or by
using conventional electrocoalescence.
The paper presents a methodology that permits the optimization of water and oil separation in the case of an in situ extra-
heavy oil (produced by thermal process). The crude oil was first characterized in terms of rheological behavior and interfacial
properties. The dilatational viscoelastic properties of the interface were determined from measurements performed with an
oscillating oil drop tensiometer. Properties of emulsification were also investigated by using a specific device named
"dispersion rig" that allows the reconstitution of crude oil emulsions under controlled hydrodynamic conditions. Then a
laboratory procedure based on electrical stability tests was applied in order to optimize the concentration of demulsifier
required for effective water separation.
Finally, the optimal electrical parameters were determined in an electrocoalescer device in presence of the selected
concentration of additive. The efficiency of coalescence was measured by following the growth of dispersed water droplets
inside the emulsion using Differential Scanning Calorimetry (DSC).

Introduction
In situ extra heavy oils produced by cold or thermal methods (like e.g. Steam Assisted Gravity Drainage SAGD process) tend
to form tight and stable emulsions containing oil, water, diluent and solids. These emulsions have to be treated either using
conventional gravity-based vessels operating at high temperature, long retention times and huge chemical injection or using
more advanced technologies, such as electrostatic coalescers (1- 4).
For thermal in-situ extra heavy oils, like the ones in Athabasca, electrocoalescence is scarcely used. This paper will present
the specific case of one of these extra heavy oils for which a rigorous methodology was applied in the laboratory to optimize
water and oil separation by electrocoalescence.
First, some chemical and physico-chemical properties such as rheological behavior, interfacial and emulsification
properties were investigated. Then a laboratory procedure based on electrical stability tests was applied in order to select an
efficient demulsifier and to determine the optimal concentration required for an effective crude dehydration. Finally, emulsions
from the extra heavy crude oils were submitted to electrocoalescence experiments at high frequency on a tubular
electrocoalescer device developed in IFP. The efficiency of coalescence determined by DSC (Differential Scanning
Calorimetry) allowed to evaluate the influence of parameters like additive concentration, residence time under electrical field
and temperature.
2 SPE 121669

Materials and Methods

Fluids and additives


Crude oil EHO is an extra heavy crude oil issued from thermal production and provided by the operators directly from the field
without any emulsion breaker additive. The dehydrated crude oil properties are given in table 1. The samples were received as
water-in-oil emulsions containing around 40% of dispersed water. These emulsion samples were diluted with naphtha in order
to respect a fixed weight ratio of X% dehydrated crude oil/Y% naphtha. Then the diluted emulsions were centrifuged for 30
min at 10 000 rpm in order to separate the water phase. After centrifugation, the supernatant corresponding to X% dehydrated
crude oil/Y % naphtha was collected for the different experiments. The reconstituted w/o emulsions were prepared by using a
synthetic brine with 5 g/L of NaCl. Two different types of naphtha solvent were used: naphtha P (density @15°C: 0.7929) and
naphtha J (density @ 15°C: 0.6814).
Six emulsion breaker additives coming from different fields (cold and thermal production) were studied.

Table 1: Properties of EHO crude oil

Crude oil EHO (SAGD)


Specific gravity - kg/m3 1009.7
API° 8.6
Asphaltenes - % w 11.6 / 17.6
solvent C7/C5
Asphaltenes - % w 7.4
solvent C6 - (50/50 diluted Naphtha J)
TAN – mg KOH/g 3.3

Densities and viscosities


Densities were determined by using an Anton Paar DMA 5000 densimeter. The instrument used for rheological measurements
was a Haake RS150, which is a rotary rheometer. For flow curves determination, shear rate is fixed and shear stress is
measured under coaxial cylinder geometry at controlled temperature. Viscosity is determined in the shear rate range between
0.1 and 1000 s-1.

Interfacial properties
A tensiometer Krüss K100 equipped with the Wilhelmy plate was used to determine static interfacial tensions between oil and
water phases. Dynamic interfacial tension measurements were performed using a drop tensiometer (Tracker from IT Concept),
which is thoroughly described by Benjamins et al. (5).
The oil drop technique is an established method for measuring interfacial tensions at non-miscible liquids interface. In the
present setup, the oil drop profile is automatically analyzed and it is necessary to consider the pressure difference ΔP between
the inside and outside of the drop. The interfacial tension is deduced using the Laplace equation:
ΔP = σ (1/R1+1/R2) [1]
where: σ = interfacial tension
R1 and R2 = main radii of curvature of the drop

The Tracker oil drop tensiometer simply consists of a light source, a cell containing the oil drop, and a CCD camera which
are aligned on an optical bench. After the drop formation using a syringe and motor for pressing down the syringe, the drop
profile is digitalized through the camera and a personal computer. The monitor is used to align, check and adjust the drop. Two
values of area, volume and surface tension are calculated and recorded every second.
This apparatus is also designed to measure the dilatational elastic modulus or viscoelastic modulus E*, by programming
sinusoidal variations of the drop area. The software can be used to plot the fluctuating interfacial tension versus the relative
variations in drop area. It is then possible to determine both the absolute value of the complex modulus E =│E*│and the phase
angle φ between the changes in interfacial tension and the changes in interfacial area. In the case of homogeneous deformation
of the entire area, the absolute value E can be calculated from the maximum changes in interfacial tension Δσ and the
interfacial area ΔA by:
E=A(Δσ/ΔA) [2]
The elastic component E', which characterizes the conservative behavior of the interface and the viscous component E",
which characterizes the dissipative interfacial phenomena are then calculated from E and φ:
E'=E.cosφ [3]
and
E"=E.sinφ [4]
SPE 121669 3

Emulsion characterization by DSC


Differential Scanning Calorimetry (DSC) may be used for emulsion characterization, and it is especially suitable for water-in-
oil emulsions. DSC is based on the measurement of heat exchanges between a sample and a reference as a function of time or
temperature. The peak area of the recorded signal is related to the heat flow generated or absorbed by the sample. The sample
is cooled at a constant scanning rate until the water droplets in the emulsion crystallize. The crystallization temperature is
correlated to the droplet sizes and decreases with them: the smaller the volume of the droplets, the lower the temperature of
crystallization. This technique may be advantageously used for emulsion characterization before and after application of an
electrical stress. The droplet size evolution is directly correlated to the evolution of the crystallization temperature and energy
peaks. The principle of this technique was thoroughly presented in various previous publications (6-8).

Emulsification properties by Dispersion Rig


A specific experimental set up has been developed in order to reconstitute water-in-crude oil emulsions under controlled
pressure and temperature conditions (9, 10). The objective is to simulate the flow of crude oil and brine through a choke-valve
by pumping both fluids through a calibrated orifice that represents a model choke-valve (Figure 1). The turbulent flow created
downstream from the orifice leads to a dispersion of water droplets in the oil phase, i.e. emulsion formation. A programmable
controller allowing a perfect synchronization of both piston pumps pilots the simultaneous injection of both fluids. A
maximum volume of 500 ml of each phase can be pumped at a fixed flow rate into the supply tubes. The third pump is a
metering pump which may inject additives in line just after emulsion formation. The fluid pressure is measured at two
locations, upstream and downstream from the orifice. This allows the measurement of the permanent pressure drop ΔPperm
induced by the orifice. The produced water/oil mixture is collected in a separator that is located above the orifice and may be
then analyzed. The entire equipment is placed inside a cupboard and can be heated up to a maximum temperature of 80°C. The
maximum pressure is 150 bar in the piston pumps and 50 bar in the separator. The fluids system produced in the separator is
analyzed by measuring the phase volume of oil, emulsion and free water. The dispersed water in the emulsion phase is
measured by Karl-Fisher titration or DSC.

Heated cupboard
Pressure control of
the reservoir
Reservoir

Glass window
Emulsion ΔP
formation
ΔP Calibrated orifice

CH4 CH4

Additives
Water pump Oil pump
pump

Water injection Additives injection Oil injection

Figure 1: Dispersion Rig experimental set up

Electrical stability tests (EST)


The method and electrical tester apparatus used correspond to the ones developed for oil-based drilling mud stability control
(API 13B-2). It was recently proposed for the development of chemical demulsifiers for treating water-in-crude oil emulsions
(11). It consists in a simple electrode dropped in the fluid with an AC voltage at 340 Hz frequency. The voltage amplitude
increases progressively at a rate of 150 Volt/s and is limited to a maximum of 12.9 kVolt/cm. The current intensity is measured
and the "critical voltage" is recorded at 61 µA intensity which corresponds to the short-circuiting condition of the apparatus.
The higher the critical voltage value, the higher the stability of the emulsion. The mixing procedure used to form the w/o
4 SPE 121669

emulsion is as follows: the total volume of emulsion is 300 mL, prepared in a beaker of 600 mL. The mixing device is an
Ultra-Turrax T25 homogenizer, equipped with a S25-18G rod. The reconstituted water (5 g/L NaCl) is added drop by drop in
the oil phase and the mixing is performed for 10 min at 24000 rpm. This procedure allows to reconstitute w/o emulsions
similar to the ones obtained in the Dispersion rig. Temperature is controlled during EST test and measurements are performed
each 5 minutes on the same prepared emulsion. After EST tests, the emulsion stability is checked by leaving the emulsion for a
24-hours period at the test temperature while monitoring the volume of free water phase.

Electrocoalescer device
A laboratory prototype was developed in order to study the coalescence phenomena of water droplets inside a crude oil
emulsion flow stressed by a sinusoidal electrical field with controlled voltage and frequency (12, 13). Electrical effects
increase the droplet size above 100µm in the Stokes law domain critical for separator tank efficiency. The prototype consists in
two concentric tubes inside which the electrical field is applied, fluids circulating upwards inside the annular space (Figure 2).
On the laboratory prototype, the operating conditions are limited to the circulation of a fluid with a viscosity below 100 mPa.s
for a maximum flow rate of 1.2 m3/h under controlled temperature up to 80°C and atmospheric pressure. Various device
geometries are possible for testing. For this study the tube length is 100 cm and the annular space is 1 cm.
The residence time under the electrical field varies according to the flow rate. A stable emulsion is formed through an
Ultra-Turrax homogenizer. The emulsion is then injected in-line at the bottom of the electrocoalescer where the electrical field
parameters as voltage and frequency are controlled. The efficiency of coalescence is determined by measuring emulsion
characteristics (droplet size, crystallization energy from DSC) before and after application of the electrostatic field.

V
T P
220 cm - 0.84l

entrefer 1 cm
100 cm - 1.28l
75 cm -0.37l

100 cm - 0.38l

pump

190cm - 0.72l

560*22 cm - 20l
heating resistance

ultra-turrax

Figure 2: Electrocoalescer device

Results and Discussion

Crude oil properties


After reception without any treatment, the crude oil sample was firstly characterized using Karl Fisher titration (for water
content), DSC and optical microscopy. As shown in Figure 3, the crude sample is a polydisperse water-in-oil emulsion
containing around 40% (w/w) of dispersed water (confirmed by DSC). The droplets sizes were found between 1 and 40 µm.
SPE 121669 5

Figure 3: Photo of the crude sample by optical microscopy

After dilution with naphtha and centrifugation, densities and viscosities were determined as a function of temperature for
both naphthas P and J (table 2 and 3).

Table 2: Densities and viscosities of diluted crude oil EHO (naphtha P)

Temperature 50% EHO/50% Naphtha P 60% EHO/40% Naphtha P 70% EHO/30% Naphtha P
(°C) density (g/cm3) viscosity (cP) density (g/cm3) viscosity (cP) density (g/cm3) viscosity (cP)
20 0.879 12 0.922 86 0.949 520
30 0.872 9 0.915 56 0.942 280
40 0.865 7 0.908 38 0.935 150
50 0.858 5 0.901 26 0.929 88

Table 3: Densities and viscosities of diluted crude oil EHO (naphtha J)

Temperature 50% EHO/50% Naphtha J


(°C) density (g/cm3) viscosity (cP)
20 0.838 6
30 0.831 5
40 0.823 4
50 0.815 3

The rheological behavior of diluted crude oil EHO was always Newtonian. Figure 4 shows the variation of the viscosity
(50% EHO/50% naphtha) as a function of the inverse of temperature expressed in K-1. The viscosity evolution follows a
classical Arrhenius law:
µ = A exp (Ea/RT) [5]
where µ is the viscosity, A is a preexponential factor depending on the type of crude oil and dilution ratio, R is the ideal
gas constant (8.314 J/K/mol), Ea is the activation energy and T is the temperature expressed in K. In the present case (50%
naphtha), the calculated activation energy was around 18 kJ/mol for naphtha J and 22 kJ/mol for naphtha P.
6 SPE 121669

14
y = 0,0011e2718,8x
12
R2 = 0,9938

Viscosity (cP)
10
8 50% naphtha P
6 50% naphtha J
4
y = 0,0037e2171,3x
2
R2 = 0,9828
0
0,003 0,0031 0,0032 0,0033 0,0034 0,0035
1/T (K-1)

-1
Figure 4: Viscosity of diluted crude oil EHO with naphtha P and J as a function of the inverse of temperature (K )

Interfacial properties

The interfacial tensions measured with the Wilhelmy plate method between crude oil EHO diluted with naphtha J or P and the
aqueous phase (5 g/L NaCl) at 30°C are given in Figure 5. It is noteworthy that measurements are reproducible and no
significant influence of the nature of the naphtha was noticed. The equilibrium is quickly reached and the interfacial tension is
around 15 mN/m.

25

20
IFT (mN/m)

15

10
50% naphtha J (1)
5 50% naphtha J (2)
50% naphtha P
0
0 1000 2000 3000 4000
Time (s)

Figure 5: Interfacial tension (IFT) of diluted crude oil EHO with water (NaCl 5 g/L) at 30°C (tests (1) and (2) are repeatability tests)

The formation of the film formed by indigenous surfactants (mainly asphaltenes) at the interface can be followed through
the evolution of the viscoelastic modulus with time. Figure 6 gives E' and E" against time at 30°C and at a frequency of 0.1 Hz
in the case of diluted EHO (50% and 40% naphtha P). The elastic modulus is always much greater than the viscous modulus,
even at very short times. We observe an increase of E' with time up to a value of 11 mN/m after 4 hours (40% naphtha) and 9
mN/m (50% naphtha). Unsurprisingly, an increase of the proportion of naphtha leads to a lower elastic modulus. Concerning
the viscous modulus, it is quite constant around 3 mN/m for both dilution ratios. As many authors have reported it in the case
of solutions of asphaltenes or crude oils, the increase of the dilatational elastic modulus (similar here to the elastic component
E') with time proves that adsorbed asphaltenes (and probably other amphiphilic molecules) tend to reorganize themselves at
the crude oil/water interface over quite long periods of time.
SPE 121669 7

12

10
E' or E" (mN/m)
8
E' 50% naphtha E" 50% naphtha
6
E' 40% naphtha E" 40% naphtha
4

0
0 2 4 6 8 10 12
Time (h)

Figure 6: Evolution of the elastic modulus E' and viscous modulus E" of the system diluted EHO (naphtha P)/water (5 g/L NaCl) at
30°C (0.1 Hz)

Recent publications (14, 15) have underlined that aged interfaces created between solutions of asphaltenes and water behave
like a bidimensional gel near its gelation point. These authors showed that for a wide range of frequencies, the log-log plot of
the dilatational elastic modulus E is a straight line while the phase angle remains constant. To interpret their results, they have
used the criteria from Winter and Chambon (cited in 14) who demonstrated that for a cross linking polymer system near its
gelation point, the loss and storage modulus are parallel at the gel point with a pulsation dependence of:

G'~G"~ωn [6]

ω being the pulsation of the measurement and n being a relaxation exponent characteristic of the viscoelastic properties of the
gel (purely elastic for n approaching 1 and purely viscous for n approaching 0). At the same time, the phase angle φ is related
to exponent n by the following relationship:

φ=nπ/2 [7]

We have performed some measurements at different frequencies (from 0.05 to 1 Hz) on aged interfaces (one night) at
30°C. For diluted crude EHO (50% naphtha), the evolution of E with the frequency follows a power law with an exponent
n=0.1859 (see Figure 7). The measured phase angle is rather constant with the frequency and around 16° (average value). It is
noteworthy that this value is very close to the phase angle calculated from equation [7]: φ=16.7°. So, it seems that the
behaviour of this diluted crude oil follows the criteria from Winter and Chambon.
8 SPE 121669

20

15
E (mN/m)
10 Phi (°)

y = 12,793x 0,1859 Power law


5
R2 = 0,98

0
0 0,2 0,4 0,6 0,8 1 1,2
frequency (Hz)

Figure 7: Evolution of E and φ with the frequency for EHO diluted with 50% naphtha P (16 hours old interface, 30°C)

Emulsification properties: Dispersion rig


Crude oil EHO diluted with naphtha P (50/50) and reconstituted water (5 g/L NaCl) were simultaneously injected in the
dispersion rig through a calibrated orifice by varying the water cuts (from 20 to 80%) and the total flow rates (from 20 to 120
L/h) at 30°C. The diameter ratio β=Do/Dp was equal to 0.125, Do being the orifice diameter (0.6 mm) and Dp being the pipe
diameter (4.8 mm). In these conditions, the values of the Reynolds number in the pipe vary from 150 to 900, while they vary
from 1200 to 7000 in the orifice. So the flow may be considered as laminar upstream the orifice and as turbulent in the orifice,
which causes the emulsification between oil and water.
The measured pressure drop ΔPperm is caused by the dissipation of kinetic energy due to the turbulent flow regime existing
downstream the orifice. Under the assumption of homogeneous isotropic turbulence, the average amount of energy ε that is
being dissipated per time and per mass unit can be estimated by the following equation:

ε=ΔPperm.Uo/ρ.Ldis [8]

where Uo is the average fluid velocity in the orifice, ρ is the continuous phase density and Ldis is the length of the zone in
which most of the kinetic energy is dissipated. For a turbulent flow in a circular pipe, the general relationship currently used to
define this length is: Ldis = 2.Dp.
After the emulsion formation step, emulsified fluids are collected. After a period of rest (24h) during which separation of
phases occurs under the only action of gravity, the amount of free water is recorded and allows to calculate the amount of
water that was emulsified in the oil phase. Figures 8 a and b present for various initial water-cuts the percentage of dispersed
water in the crude oil versus the pressure drop (figure 8a) or the total flow rate (figure 8b). As expected, the formation of an
emulsion phase is induced by the mixing energy dissipated through the orifice. The water amount in the emulsion phase
increases with the flow rate or dissipated energy whatever the initial water-cut. The average droplet size of the w/o emulsion
determined by optical microscopy analysis is between 1 and 2 µm. For water-cut lower than 60%, an energy threshold exists
above which the system tends to be totally emulsified as a w/o emulsion. For water-cuts higher than 60%, whatever the level
of energy dissipated, the oil/water system is never totally emulsified. Even at a high energy level, there is a water-cut limit
where a w/o emulsion co-exists with a free water phase ("phase inversion point"). This phenomenon is clearly evidenced in
Figure 9 which gives the percentage of water emulsified in the oil phase as a function of the initial water cut for different flow
rates.
It is noteworthy that we were able to identify the formation of multiple emulsions at the critical water cut of 60% which
may be considered here as the "inversion point". According to Sherman (16), this type of emulsion forms spontaneously
during the process of emulsification of a simple emulsion, when you are close to the inversion point. In our case, we have
observed water-in-oil-in-water emulsions (W/O/W) by analyzing by optical microscopy the free water separated from the oil
phase after an experiment at an initial water cut of 60% at a total flow rate of 80 L/h. Figure 10 shows the photo of the aqueous
phase, which contains several globules of oils with small water droplets inside. This was confirmed by the DSC analysis of the
aqueous phase. The thermograms obtained during the successive cooling and reheating of the free water phase are given in
Figure 11. Figure 11a shows the complete thermogram where we were able to identify an exothermic peak characteristic of the
freezing of free water (around -15°C). Then, it is followed by a very small peak at -40°C, characteristic of the freezing of
emulsified water droplets (around 1 µm). This small peak is significant, as it is shown in Figure 11b where we have zoomed on
this particular signal. During heating, the endothermic peak at -21°C corresponds to the eutectic melting of the solidified
aqueous phase that contains 5 g/L of NaCl. The peak around 0°C corresponds to the ice melting. So the DSC analysis of the
SPE 121669 9

separated aqueous phase clearly confirms the coexistence of finely emulsified droplets and free water, which is only possible
in a w/o/w multiple emulsion.

80 80
% dispersed water

% dispersed water
60 60

40 40

20 20

0 0
0 20 40 60 80 100 120 140 0 20 40 60 80 100 120 140
ΔP (bar) Flow rate (L/h)

WC 20 WC 30 WC 40 WC 50 WC 60 WC 80 WC 20 WC 30 WC 40 WC 50 WC 60 WC 80

Figure 8: % of emulsified water in diluted crude EHO as a function of ΔP (a, left) or total flowrate (b, right) for dispersion rig
experiments at 30°C and different initial water cuts (WC)

80

60 40 L/h
% dispersed water

60 L/h
80 L/h
40
100 L/h
120 L/h

20 Reference

0
0 10 20 30 40 50 60 70 80
Initial water cut (%)

Figure 9: % of emulsified water in diluted EHO as a function of the initial water cut for different flow rates
10 SPE 121669

Figure 10: Photo of the separated aqueous phase collected after a dispersion rig experiment at the "inversion point" (initial water cut
of 60% and total flow rate of 80 L/h)

4 0,5
freezing of
3
0,3 emulsified water
freezing of
Heat Flow (mW/g)

Heat Flow (mW/g)

2 freezing of free water


emulsified water 0,1
1

0
-0,1
-1
-0,3
-2 melting of ice

-3 -0,5
-100 -80 -60 -40 -20 0 20 40 -100 -80 -60 -40 -20 0 20 40
Temperature (°C) Temperature (°C)

Figure 11: DSC thermograms of the aqueous phase at the "inversion point"; a (left): complete DSC analysis; b (right): zoom on the
peak characteristic of the presence of emulsified water

Electrical stability tests (EST)


Emulsions were prepared according to the procedure described in the Materials and Methods section. The oil phase was 50%
crude oil EHO/50% naphtha J and the water phase was an aqueous solution at 5 g/L NaCl. Temperature was fixed at 30°C and
controlled during EST tests. Measurements were performed every 5 minutes on the same prepared emulsion during 15 min.
Considering the electrode geometry, the 200 Volt applied voltage corresponds to 1.2 kVolt/cm. After EST tests, the emulsion
stability is checked by leaving the emulsion for a 1-hour and 24-hour period at the test temperature while monitoring the
volume of free water phase.
Six formulations of demulsifiers (D1 to D6) were tested, including the formulation used on the field for gravity separation
of this crude oil (demulsifier D1).
The results on voltage variation versus time for the additive D1 at several concentrations are presented in Figure 12. From
these results, it seems that 50 ppm of concentration leads to a significant decrease of the voltage required. The comparison of
all the additives is shown in Figure 13. In this case, results are normalized and give the percentage of the voltage decrease as a
function of the additive concentration. The emulsion destabilization after EST test was evaluated by measuring the amount of
free water after 24h (Figure 14). Based on these two types of results (EST tests, emulsion destabilization after EST test), the
behavior of additives could be compared:
SPE 121669 11

• All the additives induce a voltage decrease, so they have an effect on emulsion stability (Figure 13). In terms of
critical voltage decrease, additive D2 seems the most efficient at low concentration (less than 50ppm), followed
by additives D1 and D5, the less effective being D3, D4 and D6
• When there is no additive (Figure 12), there is no effect on emulsion stability
• After 24 hours (Figure 14) the water phase is completely separated with additive D1 at 50ppm while no free water
appears in case of additives D2, D4 and D5
• Additives D3 and D6 show an intermediate behavior at 50ppm (50% voltage decrease and appearance of free
water phase)
So from these results, D1, which is the formulation used on the field (for gravity separation), appears as the most efficient
because it leads to a significant decrease of the critical voltage, and it allows a complete separation of water after the EST test.
It is noteworthy that D2, which appeared as the most promising additive from the EST tests, does not allow water separation at
50 ppm. It is certainly due to the fact that it allows a quick flocculation of water droplets in the emulsion (formation of a chain
of droplets inducing short-circuiting), but not necessarily coalescence. That is the reason why it is important to record the
evolution of the emulsion after performing the EST test.
Further tests were performed with the two products D1 and D3. The effects of the two selected additives on emulsion
stability are presented in Figure 15. The variation of the amount of separated water versus the additive concentration is shown
in two cases corresponding to the stability of the initial emulsion (noted "without EST" in graph) and emulsion at 24h after
EST tests (noted "after EST"). It is noteworthy that the additive alone is not sufficient to induce a complete destabilization of
the emulsion and that the electrical stress is necessary to cause coalescence, hence separation of the oil and water phases. It is
especially true for additive D3, for which no separation of water was observed with the bottle test at 30°C. By increasing the
temperature up to 50°C, the maximum amount of separated water is below 40% at 150 ppm. It seems therefore that the
application of the electrical stress in EST tests allows a significant decrease of the amount of demulsifier required for
coalescence of droplets in the emulsion.

0 ppm
500 25 ppm
50 ppm
400 75 ppm
100 ppm
Voltage (V)

300

200

100

0
0 2 4 6 8 10 12 14 16
Time (min)
Figure 12: EST tests on emulsion 50% EHO/50% naphtha J-WC 30%-30°C-Additive D1
12 SPE 121669

1,0
D6 D3

0,8 D4 D5
Voltage EST (%)

D1 D2
0,6

0,4

0,2

0,0
0 20 40 60 80 100 120
additive concentration (ppm)
Figure 13: EST tests at 15 min on emulsion 50% EHO/50% naphtha J-WC 30%-30°C-all additives

120
D1 D6

100
D3 D5
Separated water (%)

80 D2 D4

60

40

20

0
0 20 40 60 80 100 120
Concentration (ppm)
Figure 14: % of separated water 24 hours after EST tests on emulsion 50% EHO/50% naphtha J-WC 30%-30°C-all additives
SPE 121669 13

D1 - after EST (30°C)


120 D3 - after EST (30°C)
D1 without EST (30°C)
100 D3 without EST (50°C)
D3 without EST (30°C)
Separated water (%)

80

60

40

20

0
0 50 100 150
additive concentration (ppm)

Figure 15: % of separated water 24 hours after EST tests or without EST tests on emulsion 50% EHO/50% naphtha J-WC 30%-Different
temperatures- additives D1 and D3

Electrocoalescence efficiency in the electrocoalescer rig


From the results of the EST tests performed on the different demulsifier additives, it was decided to select additive D3 for the
experiments in the electrocoalescer rig because this formulation allows a significant separation of water after EST test while
absolutely no water is separated from bottle tests at the same temperature (30°C).
Experiments on the electrocoalescer rig were performed with a w/o emulsion formed with an in-line ultra-turrax
homogenizer from an oil phase (50% crude oil EHO/ 50% naphtha J) and a water phase (5g/l NaCl brine) at 30% water-cut in
presence of additive D3 at 1 kHz frequency. This high value of frequency has been chosen following previous studies on
various types of extra-heavy oils (12, 13). The residence time of the emulsion under electrical stress is fixed by the flow rate
between 0.05 up to 1.0 m3/h that corresponds to a time range of 90 to 4.5 sec. All experiments are performed under a laminar
flow regime. It is noteworthy that w/o emulsions formed in the electrocoalescer device are similar to the ones formed for EST
tests.
The procedure consists firstly in determining the breakdown voltage where short-circuiting starts to appear. Then,
electrocoalescence efficiency is evaluated at this breakdown voltage by measuring the droplet coalescence using mainly the
DSC technique.
The electrocoalescence efficiency Ec is defined considering the evolution of the emulsion droplet distribution given by the
DSC thermogram before and after electrical stress on the w/o emulsion flowing in the tube. As indicated on the typical DSC
thermograms presented in Figures 16a and 16b, A0 is the crystallization energy of the fine droplet peak of the initial emulsion
and Af the crystallization energy of the fine droplet peak still present in the emulsion after electrical stress. This energy is
calculated from the integration of the peak. So the efficiency of electrocoalescence Ec is given by the following equation:

Ec = 1 – (Af/A0) [9]

These Figures show that after applying the electrostatic field, there is formation of a population of bigger droplet sizes
compared to the initial one, i.e. that coalescence of droplets did occur.
14 SPE 121669

1,5 1,5
freezing of
1 freezing of 1 large droplets
A0 small droplets Af
Heat Flow (mW/g)

Heat Flow (mW/g)


0,5 0,5

0 0

-0,5 -0,5
melting of
-1 ice -1

-1,5 -1,5
-100 -80 -60 -40 -20 0 20 40 -100 -80 -60 -40 -20 0 20 40
Temperature (°C) Temperature (°C)

Figure 16: Typical DSC thermograms of the emulsion before (Figure 16a, left) and after (Figure 16b, right) the application of the
electrical field in the electrocoalescer rig

For the tested emulsion of diluted crude oil in presence of additive, breakdown voltages vary between 1.8 up to 2.4 kVolt
leading to a power around 50 Watts.
Figure 17 shows the efficiency of electrocoalescence measured by DSC as a function of the D3 additive concentration for a
residence time of 45 sec at 30°C. As expected, the efficiency increases with the additive concentration. No effect of the
additive is noticed up to 25 ppm, while a concentration of 50 ppm gives an efficiency of elctrocoalescence of around 45%.
More experiments were performed at this concentration of 50 ppm of D3 by varying the residence time under electrical
field and one test was done at a temperature of 50°C at a residence time of 4.5 s. The electrocoalescence efficiency measured
from DSC analysis as well as the amount of free water separated 24 hours after the electrical field application are reported in
Figures 18a and b. Both values increase continuously with the residence time. It is noteworthy that there is a good correlation
between the efficiency of electrocoalescence and the percentage of water separated 24 h after the electrical stress. At 50 ppm,
an electrocoalescence efficiency of 45% corresponds to around 60% of separated water. It is also interesting to observe that the
temperature strongly enhances the efficiency of the process at very short time of residence (4.5 s). An increase of the
temperature up to 50°C gives an efficiency of electrocoalescence of 61% (17% at 30°C) and a percentage of free separated
water after 24h of 86% (10% at 30°C).
From these tests performed on the electrocoalescer rig, we can conclude that some of the w/o emulsions formed from
diluted crude oil EHO could be more easily broken under electrical stress with a relatively low concentration of additive D3
(50 ppm) compared to gravity separation which requires a much higher additive concentration (> 150 ppm).
SPE 121669 15

80
70
60
Efficiency (%)

50
40
30
20
10
0
0 10 20 30 40 50 60
additive concentration (ppm)

Figure 17: Efficiency of electrocoalescence determined from DSC analysis as a function of additive D3 concentration at 30°C

100 100
50°C - 50 ppm 50°C - 50ppm
80 30°C - 50 ppm 30°C - 50ppm
Separated Water (%)

80
Efficiency (%)

60 60

40 40

20 20

0 0
0 10 20 30 40 50 0 10 20 30 40 50
Residence Time (s) Residence Time (s)

Figures 18: Efficiency of electrocolaescence (a, left) and percentage of separated water after 24h (b, right) as a function of residence
time under electrical field for 50 ppm of D3, at 30 and 50°C

Conclusion

From this study, it is clear that stable w/o emulsions form easily from extra heavy oil EHO. Dispersion rig experiments
show that this oil, diluted with naphtha to reach a viscosity around 5-10 cP at 30°C, is able to incorporate up to 60% of water
at high energy levels. The formed emulsions are very fine (average diameter around 2 µm) and are stabilized by polar heavy
components, such as asphaltenes, structured as a rigid film at the water/oil interface, as highlighted by the interfacial
measurements performed with the oscillating drop apparatus. The measured viscoelastic properties of the w/o interface show
that the elastic modulus is much greater than the viscous modulus, even at very short times. Furthermore, the aged interface
created between diluted EHO and water seems to behave like a bidimensional gel near its gelation point, as it was recently
underlined by other authors on solutions of asphaltenes or diluted crude oils. These interfacial properties consequently limit
the coalescence phenomena and hence limit the efficiency of separation by gravity.
Bottle tests performed with formulation of demulsifiers confirm that it is very difficult to destabilize these emulsions by the
only action of gravity. Tests performed with the Electrical Stability Tester (EST) allow to select demulsifiers showing an effect
on emulsion stability. However, the critical voltage measured by this way does not give direct information about the
mechanism of action of additives, i.e. flocculation or coalescence or both. It is therefore necessary to complete the test by the
analysis of the emulsion after the application of the electrical stress, by recording the amount of free separated water, which is
the consequence of en effective coalescence between water droplets. All the tests performed according to this procedure tend
to demonstrate that the amount of demulsifier additive required for separation of oil and water by electrocoalescence is much
lower than for gravity separation.
16 SPE 121669

This tendency was clearly confirmed by the tests achieved in the electrocoalescer rig set up. High electrical stress (high
voltage, high frequency) during short duration (< 60 sec) and in combination with a chemical demulsifier leads to some water
droplets coalescence. Each stress by itself (electrical or chemical) is not efficient enough for emulsion breakage for this type of
oil.
So the methodology proposed here may be advantageously used to characterize and compare emulsions, select
formulations of additive for dehydration of extra-heavy oils by electrocoalescence, keeping in mind that the best demulsifiers
are not necessarily the same depending on the type of dehydration process (gravity separation or electrocoalescence).

Acknowledgement
The authors wish to thank IFP and TOTAL for the permission to publish and A. Mouret, P. Thoral and L. Podesta-Foley for
the experimental work.

References
1. SAMS G.W. and ZAOUK M., Emulsion resolution in electrostatic processes, Energy & Fuels, 2000, 14, 31-37.
2. NOIK C., CHEN J., and DALMAZZONE C., Electrostatic demulsification on crude oil: A state-of-the-art review, SPE
International Oil & Gas Conference and Exhibition in China, 5-7 December 2006, Beijing, China, SPE 103808-PP.
3. EOW J. S., GHADIRI M., Electrostatic enhancement of coalescence of water droplets in oil: A review of the technology,
Chemical Engineering Journal, 2002, 85, 357-368.
4. NOIK, C., DALMAZZONE, C., GOULAY, C., and GLENAT, P., Characterisation and emulsion behaviour of Athabasca extra
heavy oil produced by SAGD, SPE/PS-CIM/CHOA SPE Int.Thermal Operations and Heavy Oil Symp. held in Calgary, Alberta,
Canada, 1-3 November 2005, SPE-97748.
5. BENJAMINS, J., CAGNA, A., LUCASSEN-REYNDERS, E.H., Viscoelastic Properties of Triacylglycerol/Water Interfaces
Covered by Proteins, Colloids and Surfaces A, 1996, 114, 245-254.
6. DALMAZZONE, C., and CLAUSSE, D., Microcalorimetry, Ch. 14 In: J Sjöblom, ed., Encyclopedic Handbook of Emulsion
Technology, Marcel Dekker, New York, 2001, 327-347.
7. CLAUSSE, D., GOMEZ, F., PEZRON, I., KOMUNJER, L., and DALMAZZONE, C., Morphology characterization of emulsions
by differential scanning calorimetry, Advances in Colloid and Interface Science, 2005, 117, 59-74.
8. CLAUSSE, D., GOMEZ, F., DALMAZZONE, C., NOIK, C., A method for the characterization of emulsions:
Thermogranulometry. Application to water-in-crude oil emulsion, Journal of Colloid and Interface Science, 2005, 287, 694-703.
9. MALOT, H., NOIK, C. and DALMAZZONE, C., Experimental investigation on water-in-crude oil emulsion formation through a
model choke-valve: droplet break-up and phase inversion. In Proceedings Multiphase 2003 Conference, 11-13 June 2003, San
Remo (Italy), BHR Group 2003 Multiphase Technology, 543-559.
10. GALINAT, S., MASBERNAT, O., GUIRAUD P., DALMAZZONE C. and NOIK, C., Drop break-up in turbulent pipe flow
downstream of a restriction, Chemical Engineering Science, 2005, 60, n°23, 6511-6528.
11. BEETGE, J.H., HORNE, B.O., Chemical demulsifier development based on critical electric field measurements, SPE International
Symposium on Oilfield Chemistry, 2-4 February 2005, Houston, Tx, USA, SPE 93325.
12. NOIK, C., DALMAZZONE, C., GLENAT, P., PICHOT, D., Extra Heavy Oil pre-Dehydration Process Development, World
Heavy Oil Congress, 10-12 March 2008, Edmonton, Alberta, Canada, Paper 2008-319.
13. NOIK, C., DALMAZZONE, C., GLENAT, P., Pre-electrocoalescer unit adapted to the extra-heavy oils characteristics, 2008 SPE
International Thermal Operations and Heavy Oil Symposium, 20-23 October 2008, Calgary, Alberta, Canada, SPE 117563.
14. BOURIAT, P., EL KERRI, N., GRACIAA, A., LACHAISE, J., Properties of a Two-Dimensional Asphaltene Network at the
Water-Cyclohexane Interface Deduced from Dynamic Tensiometry, Langmuir, 2004, 20, 7459-7464.
15. DICHARRY, C., ARLA, D., SINQUIN, A., GRACIAA, A., BOURIAT, P., Stability of Water/Crude Oil Emulsions based on
Interfacial Dilatational Rheology, Journal of Colloid and Interface Science, 2006, 297, 785-791.
16. SHERMAN, P., Emulsion Science, P. Sherman (ed.) Academic Press: London, 1986, 206-207.

Você também pode gostar