Вы находитесь на странице: 1из 15

This article was downloaded by: [L'Oreal]

On: 10 March 2014, At: 12:10


Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered office: Mortimer House,
37-41 Mortimer Street, London W1T 3JH, UK
Journal of Dispersion Science and Technology
Publication details, including instructions for authors and subscription information:
http://www.tandfonline.com/loi/ldis20
Emulsions of Heavy Crude Oils. II. Viscous
Responses and Their Influence on Emulsion Stability
Measurements
Anne Silset
a
, Andreas Hannisdal
b
, Pl Viggo Hemmingsen
c
& Johan Sjblom
a
a
Ugelstad Laboratory, Department of Chemical Engineering , Norwegian University of
Science and Technology (NTNU) , Trondheim, Norway
b
Aibel Technology , Billingstad, Norway
c
StatoilHydro, Research Centre , Trondheim, Norway
Published online: 17 Sep 2010.
To cite this article: Anne Silset , Andreas Hannisdal , Pl Viggo Hemmingsen & Johan Sjblom (2010) Emulsions of Heavy
Crude Oils. II. Viscous Responses and Their Influence on Emulsion Stability Measurements, Journal of Dispersion Science and
Technology, 31:10, 1432-1445, DOI: 10.1080/01932690903210341
To link to this article: http://dx.doi.org/10.1080/01932690903210341
PLEASE SCROLL DOWN FOR ARTICLE
Taylor & Francis makes every effort to ensure the accuracy of all the information (the Content) contained
in the publications on our platform. However, Taylor & Francis, our agents, and our licensors make no
representations or warranties whatsoever as to the accuracy, completeness, or suitability for any purpose of the
Content. Any opinions and views expressed in this publication are the opinions and views of the authors, and
are not the views of or endorsed by Taylor & Francis. The accuracy of the Content should not be relied upon and
should be independently verified with primary sources of information. Taylor and Francis shall not be liable for
any losses, actions, claims, proceedings, demands, costs, expenses, damages, and other liabilities whatsoever
or howsoever caused arising directly or indirectly in connection with, in relation to or arising out of the use of
the Content.
This article may be used for research, teaching, and private study purposes. Any substantial or systematic
reproduction, redistribution, reselling, loan, sub-licensing, systematic supply, or distribution in any
form to anyone is expressly forbidden. Terms & Conditions of access and use can be found at http://
www.tandfonline.com/page/terms-and-conditions
Emulsions of Heavy Crude Oils. II. Viscous Responses
and Their Inuence on Emulsion Stability Measurements
Anne Silset,
1
Andreas Hannisdal,
2
Pal Viggo Hemmingsen,
3
and
Johan Sjoblom
1
1
Ugelstad Laboratory, Department of Chemical Engineering, Norwegian University of Science and
Technology (NTNU), Trondheim, Norway
2
Aibel Technology, Billingstad, Norway
3
StatoilHydro, Research Centre, Trondheim, Norway
The stability of 30 heavy crude oil emulsions was studied in a parallel-plate laboratory coalescer
(DC eld). Particularly, viscous responses and their inuence on the emulsion stability measure-
ments were investigated. In addition to highlighting previous results from the same experimental
setup and discussing these based on recent experience, new results at different temperatures and
volume fractions of water were presented. A new semi-empirical model for the characteristic time
of the destabilization process was presented. The electrical forces were modelled with a
point-dipole approximation and the hydrodynamic resistance to droplet transport was modelled
with an empirical term including the logarithmic viscosity of the oil phase. The new model clearly
performed much better than the previous model, particularly for very viscous crude oils. Studies of
the performance of industrial electrocoalescers have showed that simple electrostatic theory can
potentially explain complex separation phenomena when the resistance to the coalescence step is
reduced by an efcient demulsier. The ultimate goal is to build a model for both the laboratory
setup and the industrial coalescer so that laboratory experiments can be used to predict the
behavior of the industrial process.
Keywords Crude oil, electrocoalescence, emulsion stability, point-dipole approximation,
temperature, viscosity, water cut
1. INTRODUCTION
In the petroleum industry, stable water-in-oil (w=o)
emulsions will be formed when oil and coproduced water
is mixed in the reservoir during transportation from the
wellhead to the platform or in the process plant. The water
has to be removed before rening takes place. It is very
important to be aware of the properties that inuence the
formation of emulsionsas well as the destabilization
mechanisms of the emulsion system. The destabilization
mechanism of an emulsion involves two general steps
occulation and coalescence. Depending on the method-
ology of separation, the occulation step is at least
controlled by factors such as the density and viscosity of
the continuous oil phase and the sizes of the dispersed
water droplets. The coalescence step is controlled by
natural surfactants within the crude oil and indigenous
compounds like naphthenic acids, resins, and asphaltenes,
which stabilize the interfacial lm on the water droplets.
Inorganic solids, made interfacially active by adsorption
of organic molecules, can also contribute considerably to
the overall stability of the interfaces. The interfacial
chemistry of petroleum w=o emulsions has been studied
in detail previously.
[15]
Several techniques exist for enhancing the separation of
w=o emulsions in industrial processes, such as chemical,
[6,7]
gravitational, or centrifugal settling,
[8]
heat treatment and
electrostatic destabilization.
[9,10]
In the petroleum industry,
the electrocoalescence separation concept has been used
both for water knock-out processes and polishing pro-
cesses. Strong electrical elds create attractive forces
between conductive water droplets (with salt), which are
dispersed in the insulating crude oil. The forces enhance
both the occulation and the coalescence step in the
dehydration process.
[11,12]
Cottrell and Speed
[11]
led the rst patent on electrocoa-
lescence, observing the coalescence mechanism when a high
potential was applied to a pair of wire electrodes in a w=o
Received 3 December 2008; accepted 29 December 2008.
The authors would like to acknowledge the nancial support
from the participants in the Joint Industry Project, Flucha III,
hosted by the Ugelstad Laboratory at the Norwegian University
of Science and Technology.
Address correspondence to Andreas Hannisdal, Aibel
Technology, Bergerveien 12, 1396, Billingstad, Norway. E-mail:
Andreas.Hannisdal@aibeltechnology.com
Journal of Dispersion Science and Technology, 31:14321445, 2010
Copyright # Taylor & Francis Group, LLC
ISSN: 0193-2691 print=1532-2351 online
DOI: 10.1080/01932690903210341
1432
D
o
w
n
l
o
a
d
e
d

b
y

[
L
'
O
r
e
a
l
]

a
t

1
2
:
1
0

1
0

M
a
r
c
h

2
0
1
4

emulsion. Chains of aqueous drops extended from one
electrode to the other. Coalescence of adjacent drops in
each chain then occurred and the drops next to the electro-
des became larger as new drops were acquired by the chain.
Almost any type of high electric eld will assist the separ-
ation of w=o dispersions to some extent.
[13]
Alternating
current (AC), direct current (DC), pulsed DC, or combina-
tions of them are being utilized in the separation of
water-in-crude oil emulsions. Each eld type acts according
to different mechanisms. Most of the available commercial
electrocoalescers work in AC because of their high water
tolerance, effective power consumption, and low tendency
for electrolytic corrosion. The DC electric eld has been
less common in the past. In 1981, the concept of pulsed
DC elds was introduced, together with insulated electro-
des. Since then, this has become more common in the
electrocoalescence technology. Pulsed DC and AC elds
are especially useful, when the aqueous phase content of
the emulsion is high, to prevent short circuiting the electri-
cal system.
[14]
Generally, two main designs of coalescers are common;
cellular units, which may consist of coated or uncoated elec-
trodes and either internal or external settlers, and tank
units, with bare electrodes in which coalescence and settling
take place simultaneously, as commonly used in the oil
industry. Traditional electrocoalescers have nonisolated
electrodes and are for that reason placed downstream nor-
mal gravity separators to avoid too high water loading and
short circuiting followed by collapse of the electric eld.
Recently, other nontraditional electrostatic coalescer units
have appeared in the market, that is, the Compact Electro
Coalescer (CEC) from Aker Solutions (Oslo, Norway),
the Vessel Internal Electrostatic Coalescer (VIEC) from
Aibel (Asker, Norway), and the Dual Polarity Treater from
Natco (Houston TX, USA). All these products have
partially or completely isolated electrodes to reduce the
potential for short circuiting at high water cuts.
[15,16]
The
CEC is an inline AC coalescer, the VIEC is built as a wall
of AC coalescer elements in the rst or second stage
separator, and the Dual Polarity Treater is a compact dehy-
drator tank utilizing both DC and AC elds. The character-
istics and geometry of the electrodes and the electrical eld
determine the performance of the electrostatic coalescer.
To our knowledge, Kilpatrick and Speicker
[17]
were
among the rst to use electrical elds in a parallel-plate
capacitor to evaluate w=o emulsion stability. The experi-
mental setup was similar to the one used in this study: par-
allel nonisolated electrodes, direct current, and constantly
increasing eld strength until instability occurred. The
point of instability was identied by monitoring the current
through the emulsion. Aske et al.
[18]
used the same labora-
tory setup to investigate which analytical and physico-
chemical parameters that contributed to emulsion
stability in electrical elds. Using multivariate data
analysis, the emulsion stability was correlated to
physicochemical properties such as molecular weight,
density, viscosity, interfacial tension, interfacial elasticity,
total acid number (TAN), and SARA (saturates, aro-
matics, resins, asphaltenes) composition for a crude oil
matrix consisting of 21 samples. It was concluded that
the asphaltene content, the state of asphaltene aggregation,
and interfacial elasticity were the most important factors
with regard to emulsion stability. Chiesa et al.
[19]
performed a fundamental study on the combined effect of
viscosity and electrostatic forces on two water droplets
colliding in an electric eld. According to Chiesa et al., vis-
cosity inuenced indirectly the rest time of coalescing water
droplets since the impact velocity and consequently the
kinetic energy of colliding water droplets was dependent
on both the strength of electrical forces and on the viscosity
of the surrounding oil. Less et al.
[20]
studied the electro-
rheological behavior of petroleum emulsions. Droplet
transport and coalescence processes were studied in electric
elds. Less demonstrated the reversibility of droplet chain
formation caused by electric forces at low shear rates. This
phenomenon was not observed at high shear. Hemmingsen
et al.
[21]
studied the rheology and the stability of 27 heavy
crude oils in DC electric elds. Emulsion stability was
measured with the same parallel-plate capacitor as the
one used in the current study. The magnitude of the
electrical eld was constantly increased with a step rate,
dE
0
=dt, until destabilization occurred. Hemmingsen et al.
realized that viscous responses greatly inuenced the
reorganization of water droplets in the electric eld. Tem-
perature and dilution effects were studied. Based on the
ndings by Hemmingsen et al.,
[21]
Hannisdal et al.
[22]
extended the work and included experiments with varying
dE
0
=dt. Thirty crude oils were analyzed with respect to dif-
ferent bulk properties, interfacial properties, spectroscopic
signatures, and emulsion stability. Emulsion stability was
measured with the same parallel-plate capacitor as the
one used in the current study. The experimental results
were compared to a simplied theoretical model that
described the destabilization process by taking drag forces
and dielectrophoretic forces into account. To a large
extent, the results could be explained by the rate of droplet
transport and the viscosity of the continuous phase. Given
enough time, water-in-heavy oil emulsions could be desta-
bilized even at very low electrical eld magnitudes. Even
though water-in-heavy crude oil emulsions are expected
to have a considerable barrier to drop coalescence through
the lm strength, such interfaces seemed to provide limited
stability to electrically induced disintegration. Atten
[23]
came to the same conclusion when the performance of a
coaxial cylindrical electrocoalescer (AC eld) was evalu-
ated. Beetge and Horne
[24]
studied the electrical destabiliza-
tion of emulsions with different amount of water to nd the
relative amount of energy required for occulation and
EMULSIONS OF HEAVY CRUDE OILS 1433
D
o
w
n
l
o
a
d
e
d

b
y

[
L
'
O
r
e
a
l
]

a
t

1
2
:
1
0

1
0

M
a
r
c
h

2
0
1
4

coalescence. Crude oil blends and model oil emulsions had
in some cases a distinct barrier toward coalescence, which
diminished by introducing demulsier. Even though the
theoretical analyses and the interpretation of results differ
slightly, the observations by Beetge and Horne are in
harmony with results from the current study.
Here, emulsions of 30 crude oils, previously studied by
Hemmingsen et al.
[21]
and Hannisdal et al.
[22]
have been
investigated in a parallel-plate DC laboratory coalescer.
In addition to highlighting previous results and discuss-
ing these based on recent experience, new results are
presented. The focus is on viscous responses and their
inuence on the emulsion stability measurements in the
laboratory coalescer. A new semi-empirical model that
explains the experimental ndings is presented. The
model accounts for varying water cuts, magnitudes of
the electrical eld, and a broad range of oil viscosities,
both for different crude oils and for oils at different
temperatures.
2. THEORETICAL BACKGROUND
Electrocoalescence is governed by the effect of electro-
static forces. When an electric eld is applied to a w=o
emulsion, a suspended droplet is subjected to different
forces. One mechanism leading to coalescence of small dro-
plets is dielectrophoresis. Dielectrophoretic forces are
attractive forces established in a nonuniform electric eld
between droplets having permittivity, which differs from
the permittivity of the carrying liquid.
[23]
These forces pull
the droplet toward the highest voltage gradient and are
proportional to the droplet diameter and the oil conduc-
tivity, as shown by:
[26]
F
DP
4pr
3
e
0
e

W
e

0
e

W
2e

0

rE
2
; 2:1
where r is the droplet radius, E
o
and E
w
are the oil and water
permittivities, the subscripted symbol

indicates their
complex value (equal to E jrx, where r and x are conduc-
tivity and eld frequency), and rE is the electric eld
gradient. Another possible mechanism is electrophoresis.
Electrophoretic forces are both attractive and repulsive
forces established in a uniform voltage eld between
charged droplets and the electrodes. The fundamental prin-
ciple relies on the charge separation between the particle
surface and the surrounding uid. An applied electric eld
acts on the resulting charge density, causing the particle,
the surrounding uid or both to move.
[27]
Equation (2.2)
shows the formulation of the electrophoretic force for a
droplet charged by direct contact with an electrode at the
time t t
0
. This droplet will gradually lose its charge with
a relaxation time s (e
0
=r
0
).
[26]
F
EP

p
2
6

4pr
2
e
0
E
2
e
r
0
t
e
0

2:2
The remaining basic process is the dipolar force, which
causes a mutual attraction between neighboring droplets
due to the interaction of the dipoles induced by the electric
eld.
[23]
This force arises because of the high conductivity
of the water, which has high salt content and according
to Eow et al.
[27]
can be written as:
F
d

24pe
0
E
2
r
6
s
4
; 2:3
where s is the distance between two droplets radii. The
main equations are now written and can be analyzed.
The most obvious feature of all electrostatic forces is the
importance of the magnitude of the electric eld, E, which
appear in second order in all Equations (2.1) through (2.3).
Equation (2.3) shows that the dipolar force is strongly
dependent on the water droplets size and the spacing
between them, which is reasonable since it is the presence
of nearby droplets, which actually create the force.
Assuming that the water droplets have similar size and
are homogeneously distributed, the spacing is inversely
proportional to the dispersed water volume fraction /
w
,
as shown by:
[28]
s r
4=3p
/
w

4=3
: 2:4
Therefore, as the water content is reduced along the elec-
trocoalescence process because of the ongoing separation
(or just because the emulsion to be treated has a low water
content, but still far from export quality), the spacing
between the droplets is large and the dipolar attraction
becomes less and less effective.
Equation (2.2) does not show the interaction of drop
pairs, but it can be mentioned that the interaction force
follows the law by Coulomb, who showed that the force
between two point charges falls inversely with the square
of the distance between them. Under DC electric elds,
because of the unchanging polarity, this force will induce
the charged droplets to migrate toward the oppositely
charged electrode in a continuous path with a velocity
determined by the viscosity of the continuous phase. In
doing so, it is likely that droplets will collide with each
other, merging together.
[29]
In AC elds no net charge is
imparted to the droplets, so they will just oscillate around
a mean position at a frequency twice that of the electric
eld
[26]
and they will coalesce only by polarization effects.
It has been shown that electrophoresis is by far the
strongest electrical mechanism available.
[30]
Unfortunately,
the corrosion mechanisms promoted by the unidirectional
currents and the poor water tolerance have discouraged
the use of DC elds.
[27]
1434 A. SILSET ET AL.
D
o
w
n
l
o
a
d
e
d

b
y

[
L
'
O
r
e
a
l
]

a
t

1
2
:
1
0

1
0

M
a
r
c
h

2
0
1
4

Equation (2.1) points out that the dielectrophoretic
force is still independent of the droplet spacing, but it
depends on the gradient of the electric eld. Dielectrophor-
esis does not require charged particles and arises from the
fact that any dipole has a nite separation of equal
amounts of positive and negative charges. When a nonuni-
form eld is in alignment with the dipole, one end of the
dipole will be in a weaker eld, resulting in a net force pull-
ing the dipole toward the place of greatest eld intensity.
Dielectrophoresis is the key factor in dipole coalescence,
where it can be thought that a pair of close water drops
in oil attracts each other, both trying to reach the point
of maximum eld intensity between them.
[27]
Although die-
lectrophoresis requires nonuniform electric elds, it exists
in uniform elds as well, since in practice the presence of
several droplets always acts to distort the eld lines.
Having much higher permittivity than the oil, the dielectro-
phoretic force will pull the water droplets towards regions
of higher eld strength (positive electrophoresis).
[26]
3. EXPERIMENTAL
3.1. Materials
Thirty dead crude oils were received from different oil
companies with production sites on the Norwegian Conti-
nental Shelf, in the South China Sea, the Gulf of Mexico,
the UK Continental Shelf, France, Brazil, West Africa,
and Alaska, USA. About half of the oils are dened as
heavy (<API 20), while the lightest sample has an API
gravity of 33.6. Some crude oils are extremely viscous
(see Appendix 1). The crude oils have systematically been
analyzed with regard to viscosity, density, SARA fraction-
ation, TAN, interfacial elasticity, and interfacial tension
(IFT).
[22]
(Some characteristic properties are attached in
Appendix 1.) Because of the great variety in crude oil
properties, the operators meet different challenges with
respect to ow assurance and quality of the well product
due to coproduced water solids, asphaltene precipitation
or high viscosity of the oil phase. Some of the samples also
contain wax, which can be seen from their Bingham plastic
behavior below 25

C. The wax content has been investi-


gated by Hemmingsen et al.
[21]
All crude oils were sampled
in a way that minimized contamination by production
chemicals. Some samples were received with considerable
amount of coproduced water, which was removed by grav-
itationally induced separation. For some samples with high
viscosity, the residual water could not be removed easily
and the w=o systems were analyzed as received.
A model oil system was used to study the effect of vary-
ing the volume fraction of water in emulsions destabilized
by electrical forces. The model oil was made from an
aliphatic solvent with viscosity of 1.8 mPas (Telasol) and
1 vol% sorbitan monooleate.
3.2. Destabilization of Emulsions in an Electrical Field:
The E-Critical Cell
The critical electric eld cell, was used to investigate the
w=o emulsion stability (Figure 1). The technique has been
used and described previously.
[21,22,3234]
The cell is made
up of a Teon plate with a hole in the center (r 5 mm),
and a brass plate on each side. The distance between the
plates varied from 0.25 mm up to 1.0 mm. The brass plates
were connected to a computer-controlled power supply
(Agilent Model 6634B; Agilent, Santa Clara, CA, USA)
that can deliver a maximum of 100 V DC. Electrodes with
parallel-plate geometry were used. The cell was placed in a
heating cabinet to control the temperature. The crude oil
samples were emulsied at 40

C and 60

C with an Ultra
Turrax T18 basic rotor-stator homogenizer (IKA; Staufen,
Germany) at 24000 rpm for 2 minutes. The emulsion
samples were then injected into the cell. An increasing elec-
tric eld magnitude was applied over the emulsion (starting
from 0 V). Practical considerations (low voltage power
supply because of safety issues, 0100 V) limited the eld
strength in the test to 4.0 kV=cm. The current that passed
through the sample was continuously measured. The criti-
cal electric eld (CEF) was dened as the electric eld
necessary to achieve a sudden increase in the current
through the emulsion, caused by a short circuiting of the
eld. Several parallels were performed for each sample to
ensure reproducible results.
3.3. Destabilization of Emulsions in an AC
Electrical Field
Model oil emulsions with different volume fraction of
water were studied to highlight the behavior of extremely
viscous emulsions. The setup consisted of a frequency gen-
erator (1.8 kHz), and an amplier providing 1.5 kV=cm AC
across two curved, rectangular electrodes. A glass bottle
tube (100 ml) with the emulsion was placed between the
two electrodes and served as electrical insulation from the
electrodes. The setup was presented schematically by Less
et al.
[33]
FIG. 1. The gure shows the critical electric eld cell. Two brass
electrodes are connected to a power supply (DC) and an ampere meter.
EMULSIONS OF HEAVY CRUDE OILS 1435
D
o
w
n
l
o
a
d
e
d

b
y

[
L
'
O
r
e
a
l
]

a
t

1
2
:
1
0

1
0

M
a
r
c
h

2
0
1
4

4. RESULTS AND DISCUSSION
4.1. Characteristics of the Electrical Field in the
E-Critical Cell
Emulsion stability to electrically induced separation was
tested with the critical electric eld cell with a homogeneous
background DC eld increasing linearity in magnitude with
time. It is important to seek an understanding of the experi-
mental observations and to couple these to theory. The elec-
trostatic theory in electrocoalescence processes has already
been presented in previous sections. It is clear that the elec-
trode conguration and the characteristics of the applied
electric eld greatly inuences the forces present in the coa-
lescer or cell. Because of noninsulated electrodes and DC
elds, the critical electric eld cell can potentially give elec-
trophoretic mobility to the emulsied water droplets. This
requires water droplets to be charged, either by indigenous
crude oil components or by direct contact with an electrode.
Charged water droplets will migrate toward the oppositely
charged electrode resulting in collision with oppositely
charged droplets. Dielectrophoretic forces will certainly be
present in the critical electric eld cell at small droplet-droplet
distances and result in droplet attraction, chain-formation,
and nally coalescence. Dipole-dipole interactions will be
explained by Equation (2.3) at large inter-droplet distances
(s=r >>1). It seems like the dominance of either forces will
depend on the probability of droplet-electrode coalescence
and on the relaxation time of a charged droplet. It would
be highly informative to include visual observation of the
phenomena occurring in the critical electric eld cell. How-
ever, even though some interesting phenomena have been
observed for low water cut emulsion (<1%) and with a very
high resolution near-infrared camera, observation in high
water cut crude oil emulsions are limited by severe light-
scattering. Other information will be used to discuss probable
destabilization mechanisms.
Figure 2 illustrates how emulsions behave under the inu-
ence of the applied eld: a) When no eld was applied, the
water droplets were distributed mainly as single droplets;
b) as the eld increased, the water droplets moved or coa-
lesced to a rearrangement that resulted in an increase of cur-
rent through the electrical circuit; nally, the applied electric
eld was so high that the droplets coalesced (inter-droplet
lms were thinned to critical thickness), resulting in a
water-continuous bridge between the electrodes. The con-
ductive water phase (3.5 wt% NaCl) short circuited the elec-
trical circuit. The emulsion stability was dened by the
magnitude of the electrical eld when the current through
the electrical circuit showed a sudden rise.
[32]
The gradual
increase in current through the sample agrees well with the
dipolar and dielectrophoretic rearrangement mechanism
where droplets, due to polarization, line up as chains along
electrical eld lines (between the electrodes). The chains
eventually bridge the two electrodes at the critical electric
eld value. Less et al.
[20]
presented some interesting nd-
ings that can give additional insight into the destabiliza-
tion mechanism in the critical electric eld cell. Even
though Less et al. studied crude oil emulsions in a con-
centric cell with an additional shear component, the
experimental setup was very similar to the critical electric
eld cell. Both setups had noninsulated electrodes
(separated by less than 0.6 mm) and used a DC eld.
As mentioned already, Less et al. also studied the effect
of shear by rotating one of the electrodes and measuring
the resulting force (standard rheometer). Two curves
have been extracted with permission from Less et al.
Figure 3 shows continuous viscosity measurements during
a critical electric eld test. The electric eld was increased
with a constant rate as in the current study. For low shear
experiments, this resulted in a clear increase in the viscosity,
followed by a sudden irreversible decrease in viscosity.
High shear experiments did not show the maximum in vis-
cosity. Less et al. explained the observation with droplet
rearrangement and chain formation at low shear, followed
by coalescence. Droplet rearrangement is the only mech-
anism that can give an increase in viscosity as any droplet
growth with a constant water volume fraction will result
in reduced emulsion viscosity. Interestingly, Less et al. also
proved that the rearrangement mechanism was reversible.
When the electrical eld was turned off prior to reaching
a maximum in viscosity, the emulsion returned to its initial
character. All observations made by Less et al. indicated
that dipolar forces and dielectrophoresis dominated the
behavior of the emulsions.
FIG. 2. The graph shows the current through the CEF cell with an oil
continuous emulsion subjected to an increasing dc electrical eld. The
CEF value corresponds to the point where the electrodes are connected
with a bridge of conductive water droplets. The photos show the structure
of a model oil emulsion in an electric eld.
1436 A. SILSET ET AL.
D
o
w
n
l
o
a
d
e
d

b
y

[
L
'
O
r
e
a
l
]

a
t

1
2
:
1
0

1
0

M
a
r
c
h

2
0
1
4

The practical importance of electrophoresis is depending
on the relaxation time of the charged droplet after being in
contact with a noninsulated electrode. The charge will leak
to the surrounding medium with a relaxation time s (e
0
=
r
0
).
[26]
The crude oil matrix studied here has an average
conductivity (r) of 3.7 10
8
S=m (variance 7.8 10
16
)
at 40

C and a dielectric constant (e) of 2.52 (variance


0.06). This corresponds to a relaxation time constant of
about 6.0 10
4
seconds. The charged droplets will quickly
loose their charge. It seems reasonable to assume that dipole
attraction and dielectrophoresis is the dominating mech-
anism of electrocoalescence in the critical electric eld cell,
even though the experimental setup has the potential for
electrophoretic forces also.
4.2. Viscous ResponsesThe Effect of
Increasing dE
0
/dt
The effect of an electric eld and the rate of which the
eld magnitude was increased (dE
0
=dt) has been investi-
gated. Previously, the motivation for applying a constantly
increasing electric eld was to identify what was assumed
to be a critical electric eld for droplet-droplet lm rupture
and coalescence. Consequently, the experimental method
got its name: the E-critical, or the Ecrit method.
[17,31]
Moreover, by applying a constantly increasing electric
eld, a large range of crude oils could be covered in one
analysis.
Hannisdal et al.
[22]
presented results for the 30 crude oils
used in the current study, versus the step rate of the electric
eld (dE
0
=dt). The data has been presented in Figure 4. The
raw data is presented in Appendix 1. Each point repre-
sented 25 measurements of the same emulsion, with a
variance in the 00.25 kV=cm range (greatest for the most
stable emulsions). The droplet size distributions of the
emulsions were similar (mean size 6.2 2.2 mm). Emul-
sions of heavy crude oils were generally more stable than
emulsions of lighter crude oils. Some viscous emulsions
could not be destabilized at all, and were assigned the
maximum CEF value: 4.0 kV=cm. Moreover, the CEF
values increased progressively with increasing eld rate
which pointed in the direction of a viscous contribution
to the observed response parameter. The continuous phase
provided a hydrodynamic resistance to droplet motion
through its viscosity. Thus, with increasing eld rate, dro-
plets were given less time to respond to the electric eld
and the stability of the emulsion appeared to be greater.
The relative contribution from the coalescence step was
at its maximum for experiments at low eld rates, whereas
the relative contribution from the occulation step domi-
nated for experiments performed with high dE
0
=dt.
The results clearly show how the viscosity of the oil
phase affects the measured response parameter, the emul-
sion stability, both for different crude oils and for crude
oils analyzed differently. The CEF value is highly depen-
dent on how it is measured and will usually have a signi-
cant contribution from viscous effects. This is important to
remember when trying to explain emulsion stability by
other physicochemical properties of the system, as was
done previously.
[18,32]
The multivariate analysis can be
weakened by the strong covariance between the response
parameter (CEF), the viscosity, and other properties of
the crude oil.
FIG. 4. Electrocoalescence of 30 water-in-crude oil emulsions
(/
w
0.3 vol=vol) with a background DC eld. The ordinate axis repre-
sents the maximum electric eld magnitude without short circuiting
(CEF) as a function of the rate by which the electrical eld magnitude
was increased. The samples that are assigned the maximum eld magni-
tude (4 kV=cm) could not be destabilized. (Printed with permission from
Hannisdal et al.
[22]
)
FIG. 3. Viscosity as a function of time and corresponding voltage
application for emulsions (/
w
0.3) in low shear rate (6 s
1
) and high
shear rate (500 s
1
) experiments. Printed with permission from Less
et al.
[20]
.
EMULSIONS OF HEAVY CRUDE OILS 1437
D
o
w
n
l
o
a
d
e
d

b
y

[
L
'
O
r
e
a
l
]

a
t

1
2
:
1
0

1
0

M
a
r
c
h

2
0
1
4

Hannisdal et al.
[22]
modeled the droplet transport
process in an electric eld with a simple model, previously
presented by Atten.
[23]
The hydrodynamic resistance to
droplet motion was explained by Stokes law, whereas the
point-dipole approximation accounted for electric forces
acting on spherical droplets at large drop separations (s=
r >>1) lying on similar eld line. By also assuming a mono-
disperse emulsion and an initial repartition of the drops on
the vertices of a cubic lattice, the characteristic time for
droplet approach was modelled as:
s
theo

8g
15eE
2
0
p
6/
w

5=3
1
" #
: 4:1
With the viscosity (g), the permittivity (e
0
) of the continu-
ous phase, and the water cut (/
w
). By taking into account
that the electric eld magnitude was increased linearly with
time at a rate, dE
0
=dt, the characteristic time of droplet
approach could be expressed as:
s
theo

8g
5e

1=3
dE
0
dt

2=3
p
6/
w

5=3
1
" #
1=3
: 4:2
The theoretical time of droplet approach was compared
to the experimental time, s
exp
, which followed directly from
the experimental CEF value and the eld rate, dE
0
=dt.
Although very simplied, this approximation could capture
the most underlying effects, the drag on droplets
(point-dipoles) in an electric eld. Hannisdal et al.
[22]
ident-
ied a clear dependency between s
theo
and s
exp
for the 30
different crude oils. Figure 5 shows the experimental versus
theoretical time for experiments with a rate, dE
0
=
dt 0.004 kV=cms
1
. Hannisdal et al. showed that the
experimental times, s
exp
, were approximately 10 times
greater than the theoretical times, s
theo
, as indicated by
the broken line. The discrepancy between theoretical and
experimental times was discussed and it was showed that
also surfactant-free emulsions had characteristic times
much greater than expected from Equation (4.2). There-
fore, the large experimental times could not be explained
by a signicant contribution from a barrier to the coalesc-
ence step as may be anticipated from classical lm stabili-
zation mechanisms in petroleum emulsions. In fact, a
very low electric eld magnitude was sufcient to destabi-
lize the w=o interface. The time to break the emulsion of
crude oil no. 6 (g
oil
2269 cP, 40

C) at constant electric
eld (0.44.0 kV=cm) magnitude (Figure 6) followed
exactly the same trend as experiments performed at
increasing eld strength (Figure 5). The characteristic time
was underestimated by a factor of 10.
With new experience, we can conrm that the conclu-
sions by Hannisdal and et al.
[22]
were correct. For this
specic crude oil matrix, the underestimation of the
characteristic time is generally not a result of a barrier to
coalescence (thermodynamically stable), but a result of
reduced droplet transport. One may thus argue whether
or not critical electric eld is the proper name for the analy-
sis as this name gives expectations about a threshold eld
FIG. 5. The characteristic time of destabilization (s
exp
) of emulsions
(/
w
0.3, T40

C) in the DC eld is compared to the theoretical (s


theo
)
value of droplet approach as predicted from equation (4.2). The experi-
mental time was given by the measured CEF value and dE
o
=dt
(0.004 kV=cms
1
). (Printed with permission from Hannisdal et al.
[22]
)
FIG. 6. The experimental and theoretical times for emulsion no. 6
under the inuence of a constant background eld with magnitude from
0.4 to 4.0 kV=cm. The theoretical time is calculated from Equation (4.2).
(Printed with permission from Hannisdal et al.
[22]
)
1438 A. SILSET ET AL.
D
o
w
n
l
o
a
d
e
d

b
y

[
L
'
O
r
e
a
l
]

a
t

1
2
:
1
0

1
0

M
a
r
c
h

2
0
1
4

strength for achieving separation, which is not always the
case. For reasonably low viscous crude oil emulsions, the
discrepancy between experimental and theoretical values
can be solved simply by dening a factor bs
theo
=s
exp
.
In Figure 5, b is equal to 10. If b mainly is a result of the
specic experimental setup or the electrocoalescer
geometry, this opens up for huge possibilities when it
comes to predicting the performance of real processes from
laboratory experiments. Even though it has been proposed
that b is mainly dependent on the geometry of the test cell
and droplet transport phenomena, it should still be poss-
ible to identify crude oil emulsions with signicant barrier
to the coalescence step, for example because of rigid dro-
plets or particle-stabilized interfaces. These would appear
on the left side of the broken line in Figure 5. In fact, some
of the crude oils that are known to have nonstabilized
asphaltenes actually appear on the left side in Figure 5.
When identied as a problematic emulsion, an optimized
demulsier or inhibitor may bring the emulsion back to
the state where bulk properties dominate the overall emul-
sion stability. Beetge and Horne
[24]
have done important
work on understanding the relative contribution from the
occulation and the coalescence in CEF experiments.
Still, as apparent from the Figure 5 some of the most
viscous emulsions did not follow the same trend as the
remaining samples. These water-in-heavy oil emulsions
were destabilized sooner than expected from the behavior
of the remaining samples. Possible reasons were discussed
in detail by Hannisdal et al.
[22]
Among these, different elec-
trochemistry for crude oils was mentioned as a factor that
was not taken into account. Recent studies have concluded
that difference in crude oil permittivity in Equation (4.1) or
(4.2) cannot explain the behavior of the data set. The cor-
relation between experimental and theoretical character-
istic time of droplet approach for heavy oils would not
improve dramatically by allowing permittivity to vary
between different crude oils. The variance is too low
(5.3%). However, it should be mentioned that Equation
(4.1) and (4.2) assumes perfect dielectric properties of the
crude oil, which is not the case. The conductivity of the
crude oils may differ signicantly, especially at high tem-
peratures where the mobility of charge carriers in heavy
crude oils is high.
[3436]
The conductivity of heavy crude
oils may be several orders of magnitude greater than light
crude oils.
4.3. Viscous ResponsesThe Effect of
Increasing Temperature
The previous section highlighted the important viscous
contribution to droplet reorganization in electric elds,
and how this also inuenced the measured maximum eld
strength in the critical electric eld cell. Since Equation
(4.2) has already been tested for crude oils with different
viscosities, it was natural to change the temperature of
the same crude oil emulsions to produce different crude
oil viscosities. 14 of the crude oils in the data set were ana-
lyzed with respect to emulsion stability (CEF) at 60

C.
Figure 7 shows the maximum electric eld magnitude with-
out short circuiting (CEF) versus the rate that the electric
eld magnitude was increased for four water-in-crude oil
emulsions (oil no. 8, 12, 18, and 20) (/
w
0.3) at 40

C
and 60

C. Generally, the 60

C results followed the same


trend with respect to increasing eld rate (dE
0
=dt) as the
40

C results. This conrmed that the trends observed were


viscous responses caused by the viscosity of the oil phase.
The absolute difference in CEF between 40

C and 60

C
tests increased progressively with the magnitude of the
CEF. Figure 8 shows the experimental time of destabiliza-
tion for emulsions at 40

C and 60

C at a single eld rate


(0.008 kV=cms
1
). Even though some results deviated
slightly from the trend, the relative difference in CEF
between 40 and 60

C tests was constant for these 14 crude


oil emulsions (s
exp
(60)=s
exp
(40) 0.68).
According to Equation (4.2), when holding /
w
and e
constant, the temperature increase should reduce the time
for droplet approach by a scaling factor [g(60)=g(40)]
1=3
,
where g(60) is the crude oil viscosity at 60

C. The crude oils


had already been analyzed with respect to viscosity at dif-
ferent temperatures (Appendix 1). To get an impression
about what to expect from experimentally measured times,
the viscosity scaling factor [g(60)=g(40)]
1=3
was plotted with
reference to the viscosity at 40

C in Figure 9 (open
squares). The plot clearly indicated that the relative differ-
ence in time between 40

C and 60

C, s
theo
(60)=s
theo
(40),
changed with increasing viscosity of the oil phase. The
observation was a result of the characteristics of crude oil
FIG. 7. The gure shows the maximum electric eld magnitude with-
out short circuiting (CEF) versus the rate that the electric eld magnitude
was increased for 4 water-in-crude oil emulsions (oil no. 8, 12, 18, and 20)
(/
w
0.3) at 40

C and 60

C.
EMULSIONS OF HEAVY CRUDE OILS 1439
D
o
w
n
l
o
a
d
e
d

b
y

[
L
'
O
r
e
a
l
]

a
t

1
2
:
1
0

1
0

M
a
r
c
h

2
0
1
4

viscosity and its temperature dependence. Still, the obser-
vation was in contrast to Figure 8 where s
exp
(60)=s
exp
(40)
was more or less constant. Additionally, Hemmingsen
et al.
[21]
realized that the experimental CEF for 27 crude
oil emulsions (/
w
0.3) at different temperatures (T4

C
to 80

C) followed a linear relationship between CEF and


the logarithmic viscosity. Based on the observations made
by Hemmingsen et al., Equation (4.2) was modied with
a new term to take the inuence of viscosity for very heavy
crude oils into account.
s
theo
ln g
8
5e

1=3
dE
0
dt

2=3
p
6/
w

5=3
1
" #
1=3
: 4:3
By allowing the theoretical time of droplet approach to
depend on the logarithmic viscosity (in Pas), the viscosity
term in Equation (4.3) will respond differently than the
term in Equation (4.2), especially for very heavy oils with
high viscosity. The scaling factor ln[(g(60)]=ln[g(40)] was
plotted in Figure 9 with reference to the viscosity at 40

C
(lled squares). The new scaling factor was constant at high
viscosities, but was more sensitive for viscosities less than
10 cP. Figure 10 shows the comparison between experi-
mental and theoretical time (Equation (4.3)) of destabiliza-
tion for the results presented by Hemmingsen et al.
[21]
The
results include 93 data points from experiments where
the step rate of the electric eld was 0.002 kV=cms
1
.
The theoretical time (s
theo
) based on Equation (4.3) was
clearly very similar to the experimentally determined time
of destabilization, s
exp
. In addition, the least stable crude
oils separated faster than expected, probably because of
gravitational separation and coalescence before the test
was started. Finally, the results from the experiments in
Figure 4 were compared to the new time estimated from
Equation (4.3). The comparison is presented in Figure 11
FIG. 10. The characteristic time of destabilization (s
exp
) of emulsions
(/
w
0.3, T480

C) in the DC eld is compared to the theoretical


(s
theo
) value of droplet approach as predicted from Equation (4.3). The
experimental time was given by the measured CEF value and dE
o
=dt
(0.002 kV=cms
1
).
FIG. 8. Comparison between the experimental time of destabilization
for emulsions at 40

C and 60

C. The times were calculated directly from


the eld rate (0.008 kV=cms
1
) and the CEF values (kV=cm).
FIG. 9. Scaling factors to take viscosity for different crude oils at dif-
ferent temperatures into account when predicting the time for droplet
approach according to Equations (4.2) and (4.3). The symbol: g(40) means
the Newtonian viscosity in cP (or mPas) at 40

C.
1440 A. SILSET ET AL.
D
o
w
n
l
o
a
d
e
d

b
y

[
L
'
O
r
e
a
l
]

a
t

1
2
:
1
0

1
0

M
a
r
c
h

2
0
1
4

for experiments at dE
0
=dt 0.004 kV=cms
1
. Experiments
with different dE
0
=dt gave the same slope between the
theoretical and experimental time. This indicated that the
magnitude of the electrical eld was correctly accounted
for in Equation (4.3) and (4.2). The logarithmic viscosity
term in the semi-empirical model (Equation (4.3)) clearly
explained the data much better than Equation (4.2)
(Figure 5). For light crude oils, Equation (4.2) could safely
be used. However, for viscous samples, this model would
fail. Further work should focus on explaining why the log-
arithmic function is a good solution for explaining the
inuence of high crude oil viscosities. Factors like
the mechanical lm thinning process, the invalidity of the
dipole approximation as the inter-droplet distance reduces,
and other phenomena should be considered.
4.4. Viscous ResponsesThe Effect of Increasing the
Water Cut
Previous results presented by Hemmingsen et al.
[21]
showed that the CEF value was increased for emulsions
with lower volume of water droplets (/
w
0.1, 0.2, and
0.3). Aske et al.
[18,32]
came to the same conclusions when
studying emulsions with water volume fraction /
w
0.2
and 0.3. At lower water volume fraction, the distance
between droplets in the emulsion increased and the droplets
had to move longer distances in order to form linear chains
between the two electrodes. This is according to theory
(Equations (4.2) and (4.3)) a predicts longer characteristic
time of droplet approach for lower water volume fractions.
The droplet approach is assumed to be due to dipole-dipole
forces even though the experimental setup allows for
electrophoresis also. Therefore, as the water content is
reduced along the electrocoalescer in a commercial process,
because of the ongoing separation or just because the
emulsion to be treated has low water content, the spacing
between the droplets is large and the dipolar attraction
becomes less and less effective.
Beetge and Horne
[24]
investigated the stability of emul-
sions with different water cut in a critical electric eld cell
with alternating current (340 Hz) and a step rate dE
0
=dt
(150 V=s)=(0.150 cm) 0.94 kV=cms
1
. They observed very
strong dependency between the recorded CEF value and
the water cut, and CEF values as high as 12 kV=cm for
low amounts of water. The trend was correctly explained
by a occulation controlled mechanism of destabilization.
This is exactly in agreement with the topic of the current
paper: Viscous responses to electrical elds. The step
rate used by Beetge and Horne was several orders of mag-
nitude greater than the step rate used in this and other
studies
[18,21,22]
and must result in increasingly delayed
responses and high CEF values for low water cut, as for
high viscosity of the oil. It was proposed that the CEF
should be inversely proportional to the amount of water
in the emulsion (/
w
). Since Equations (4.2) and (4.3) state
that the CEF value (CEFs
heo
dE
0
=dt) is proportional to
[(p=6/
w
)
5=3
-1]
1=3
, not 1=/
w
, the two expressions have been
compared by using results from Beetge and Horne. The
CEF results were plotted as a function of X1=/
w
accord-
ing to Beetge and Horne, and X[(p=6/
w
)
5=3
-1]
1=3
accord-
ing to expressions by Atten.
[23]
The plot is presented in
Figure 12 for experiments in the range from 120

F to
FIG. 11. The characteristic time of destabilization (s
exp
) of emulsions
(/
w
0.3, T40

C) in the DC eld is compared to the theoretical (s


theo
)
value of droplet approach as predicted from equation (4.3). The experi-
mental time was given by the measured CEF value and dE
o
=dt
(0.004 kV=cms
1
).
FIG. 12. Critical electric eld for experiments performed at different
temperatures (120

F to 180
o
F).
[24]
The same data are plotted as a function
of X1=/
w
according to Beetge
[24]
and X[(p=6/
w
)
5=3
1]
1=3
according
to expressions by Atten.
[23]
EMULSIONS OF HEAVY CRUDE OILS 1441
D
o
w
n
l
o
a
d
e
d

b
y

[
L
'
O
r
e
a
l
]

a
t

1
2
:
1
0

1
0

M
a
r
c
h

2
0
1
4

180
o
F. To correctly explain the effect of increasing amount
of water in the emulsion, the data series should be linear.
The different position on the horizontal axis is not of
importance. Both expressions seem to predict the increase
in critical electric eld value relatively well. By looking at
the two expressions, Beetge and Horne clearly predict
larger CEF values at low amount of water than Atten.
For this reason, the dotted lines curve slightly for this
particular raw data. However, in the practical range of
water cut (150%), both expressions will perform well.
Finally, we will highlight a phenomenon that relates to
emulsions with large volume fraction of water and conse-
quently high emulsion viscosity. Equation (4.1) through
(4.3) only account for the viscosity of the continuous oil
phase. However, for high water cut emulsions with small
droplets, the viscosity of the emulsion may be so high that
it reduces the rate of droplet transport processes. Figure 13
shows the viscosity of model oil emulsions with varying
water volume fraction in the range from 0.3 to 0.7. As
expected, the emulsions had shear-thinning behavior and
showed a signicant increase in emulsion viscosity with
increasing number of droplets. The droplet size was
constant for all emulsions with a lognormal distribution
in the range from 1 to 45 mm (all samples were made from
the same /
w
0.7 emulsion). The emulsions did not separ-
ate any free water by gravity during two weeks. The same
emulsions (100 ml) were destabilized in AC electric elds
(1.8 kHz) between two insulated electrodes. The setup
was explained in the experimental section. Several experi-
ments were performed by varying the water volume
fraction, the time in the electrical eld and the magnitude
of the electric eld. From these experiments, it was possible
to determine the experimental time corresponding to the
phase separation of 90% of the emulsied water at a eld
strength of, for example, 1.5 kV=cm. Figure 14 shows the
experimental time of destabilization for emulsions with
different volume fraction of water. As predicted by theory,
the efciency of separation was increased dramatically by
increasing the volume fraction of water up to 0.1. However,
with further increase of /
w
, the efciency of separation
decreased signicantly (longer time to achieve separation).
We can only explain the phenomenon by the signicant
increase in emulsion viscosity that will also impact the
mobility of droplets. One could argue if the viscosity term
in Equations 4.1 through 4.3 should account for emulsion
viscosity instead of the continuous phase viscosity. How-
ever, due to the non-Newtonian character of emulsions, it
is difcult to dene the proper shear rate (see Figure 13).
The water cut, the droplet size, and the level of turbulence
will change as droplets coalesce and separate.
Sorbitan monooleate was used to stabilize the water
droplets (1 vol%, which is far above the critical micelle con-
centration of the surfactant). The self-aggregation of the
surfactant molecules produced much greater relative emul-
sion viscosities than what can be expected for petroleum
emulsions. However, the same phenomenon has also been
observed for real crude oil emulsions, even though to a
smaller extent. The experimental setup had insulated elec-
trodes and a constant magnitude of the imposed electrical
FIG. 13. Viscosity of model oil emulsions (aliphatic hydrocarbon with
1 vol% sorbitan monooleate and viscosity 1.8 mPas, 3.5 wt% NaCl water,
droplet size 145 mm).
FIG. 14. Time necessary to separate 90% of the water from 100 ml
model oil emulsions (aliphatic hydrocarbon with 1 vol% sorbitan mono-
oleate, 3.5 wt% NaCl water, droplet size 145 mm). The characteristics
of the electrical eld were: AC 1.8 kHz, 1.5 kV=cm. The triangles show
the experimental results whereas the full line shows the last term of Equa-
tion (4.2): constant x [(p=6/
w
)
5=3
1]
1=3
.
1442 A. SILSET ET AL.
D
o
w
n
l
o
a
d
e
d

b
y

[
L
'
O
r
e
a
l
]

a
t

1
2
:
1
0

1
0

M
a
r
c
h

2
0
1
4

eld (1.5 kV=cm). Still, it cannot be excluded that the
dielectric properties of the emulsion may have some impact
on the eld distribution and the magnitude of electric
forces.
5. CONCLUSIONS
The stability of heavy crude oil emulsions were studied
in a parallel-plate DC laboratory coalescer. Particularly,
viscous responses and their inuence on the emulsion stab-
ility measurements were investigated. In addition to high-
lighting previous results from the same experimental
setup and discussing these based on recent experience,
new results were presented. These included experiments
at different temperatures and volume fractions of water
(/
w
). A new semi-empirical model for the characteristic
time of the destabilization process was presented.
s
theo
ln g
8
5e

1=3
dE
0
dt

2=3
p
6/
w

5=3
1
" #
1=3
:
The new model clearly performed much better than the
previous model, especially for very viscous crude oils. The
logarithmic viscosity (g) of the continuous phase success-
fully predicted the behavior of 30 heavy crude oil emulsions
in DC electric elds. The magnitude of the electric eld was
constantly increased with a step rate dE
0
=dt until destabili-
zation occurred. The relative contribution from the
coalescence step was at its maximum for experiments at
low eld rates, whereas the relative contribution from the
occulation step dominated for experiments performed
for high dE
0
=dt. Experimental results at different step rates
conrmed that the term (dE
0
=dt)
2=3
properly explained
the magnitude of the electric eld in the destabilization
process.
Studies of the performance of industrial electrocoales-
cers (still yet to be published) have showed that simple
electrostatic theory can potentially explain complex separ-
ation phenomena when the resistance to the coalescence
step is reduced by an efcient demulsier. The ultimate
goal is to build a model for both the laboratory setup
and the industrial coalescer so that laboratory experi-
ments can be used to predict the behavior of the industrial
process.
REFERENCES
[1] Speicker, P.M. and Kilpatrick, P.K. (2004) Langmuir, 20:
40224032.
[2] Horvath-Szabo , G., Masliyah, J.H., Elliott, J.A.W.,
Yarranton, H.W., and Czarnecki, J. (2005) J. Colloid
Interface Sci., 283: 517.
[3] Freer, E.M. and Radke, C.J. (2004) J. Adhesion, 80:
481496.
[4] Wang, Y., Zhang, L., Sun, T., Zhao, S., and Yu, J. (2004)
J. Colloid Interface Sci., 270: 163170.
[5] Goual, L., Horvath-Szabo , G., Masliyah, H., and Xu, Z.H.
(2005) Langmuir, 21: 82788289.
[6] Mohammed, R.A., Luckham, P.F., and Taylor, S.E. (1994)
Colloids Surf. A, 83: 261271.
[7] Lissant, K.J. (1983) Demulsication: Industrial Application,
Surfactant; Science Series, 13. New York: Marcel Dekker.
[8] Sun, D., Duan, X.D., and Zhou, D. (1999) Colloids Surf. A,
150: 6975.
[9] Goto, M., Kondo, K., and Nakashio, F. (1989) J. Chem.
Eng. Jpn., 22 (4): 401406.
[10] Mohammed, R.A., Luckham, P.F., Taylor, S.E. (1993)
Colloids Surf. A, 80: 223235.
[11] Cottrell, F.G. (1911) Separating and collecting particles of
one liquid suspended in another liquid, US Patent 987.
[12] Cottrell, F.G. (1911) Process for separating and collecting
particles of one liquid suspended in another liquid, US Patent
98, 421 Edition.
[13] Bailes, P.J. (1984) Chem. Eng., 62: 3338.
[14] Eow, J.S. and Ghadiri, M. (2002) Chem. Eng., 85: 357368.
[15] www.aibel.com; Accessed September 30, 2010.
[16] www.akersolutions.com; Accessed September 30, 2010.
[17] Kilpatrick, P.K. and Speicker, P.M. (2001) In Encyclopedic
Handbook of Emulsion Technology, edited by J. Sjo blom;
New York: Marcel Dekker.
[18] Aske, N., Kallevik, H., and Sjo blom, J. (2002) J. Petrol. Sci.
Eng., 36: 117.
[19] Chiesa, M., Melheim, J.A., Hemmingsen, P.V., Hansen,
E.B., and Hestad, . (2006) Sep. Purif. Technol., 50: 267277.
[20] Less, S., Hannisdal, A., and Sjo blom, J. (2008) J. Dispersion
Sci. Technol., 29: 106114.
[21] Hemmingsen, P.V., Silset, A., Hannisdal, A., and Sjo blom, J.
J. Dispersion Sci. Technol., 26: 615.
[22] Hannisdal, A., Hemmingsen, P.V., Silset, A., and Sjoblom, J.
(2007) J. Dispersion Sci. Technol., 28 (4): 639652.
[23] Atten, P. (1993) J. Electrostatics, 30: 259270.
[24] Beetge, J.H. and Horne, B.O. (2005) Chemical Demulsier
Development Based on Critical Electric Field Measurements,
SPE 93325.
[25] Less, S. (2008) Ph.D. thesis, Norwegian University of Science
and Technology, Trondheim, Norway.
[26] Lundgaard, L.E., Ingebrigtsen, S., and Atten, P. (2005) In
Emulsion and Emulsion Stability; 2nd ed., edited by J.
Sjo blom; New York: Taylor and Francis, chap. 15, 549592.
[27] Eow, J.S., Sharif, A.O., and Williams, T.J. (2001) Chem.
Eng., 84: 173192.
[28] Devulapalli, R. and Jones, F. (1999) J. Haz. Mater., 70:
157170.
[29] Urdahl, O., Wayth, N.J., Frdedal, H., Williams, T.J., and
Bailey, A.G. (2001) In Encyclopedia Handbook of Emulsion
Technology, edited by J. Sio blom; New York: Marcel
Dekker, chap. 28, 679694.
[30] Warren, K.W. and Sams, G.W. (2003) The Roles of Chemical
Screening and Electrostatic Field Selection in Desalting;
Houston, TX: NATCO Group, Inc.
EMULSIONS OF HEAVY CRUDE OILS 1443
D
o
w
n
l
o
a
d
e
d

b
y

[
L
'
O
r
e
a
l
]

a
t

1
2
:
1
0

1
0

M
a
r
c
h

2
0
1
4

[31] Sjoblom, J., ye, G., Glomm, W.R., Hannisdal, A., Knag, M.,
Brandal, ., Ese, M.-H., Hemmingsen, P.V., Havre, T.E.,
Oschmann, H.-J., and Kallevik, H. (2005) In Emulsions and
Emulsion Stability, edited by J. Sjo blom; New York: Taylor
and Francis, 415476.
[32] Aske, N., Kallevik, H., and Sjo blom, J. (2002) J. Dispersion
Sci. Technol., 36: 117.
[33] Less, S., Hannisdal, A., Bjrklund, E., and Sjo blom, J. (2008)
Fuels, 87 (12): 25722581.
[34] Fotland, P. and Anndsen, H. (1996) Petr. Sci. Technol., 14
(12): 101115.
[35] Chow, R.S., Tse, D.L., and Takamura, K. (2004) Can. J. of
Eng., 82: 840845.
[36] Fotland, P., Anndsen, H., and Fadnes, F.H. (1993) Fluid
Phase Equilibria, 82: 157164.
[37] Hannisdal, A., Orr, R., and Sjo blom, J. (2007) J. Dispersion
Sci. Technol., 28: 113.
1444 A. SILSET ET AL.
D
o
w
n
l
o
a
d
e
d

b
y

[
L
'
O
r
e
a
l
]

a
t

1
2
:
1
0

1
0

M
a
r
c
h

2
0
1
4

A
P
P
E
N
D
I
X
1
C
r
u
d
e
o
i
l
c
h
a
r
a
c
t
e
r
i
s
t
i
c
s
a
n
d
r
a
w
d
a
t
a
f
r
o
m
C
E
F
e
x
p
e
r
i
m
e
n
t
s
C
h
e
m
i
c
a
l
c
o
m
p
o
s
i
t
i
o
n
a
B
u
l
k
p
r
o
p
e
r
t
i
e
s
I
n
t
e
r
f
a
c
i
a
l
p
r
o
p
e
r
t
i
e
s
b
C
E
F
a
t
4
0

C
m
e
a
s
u
r
e
d
w
i
t
h
d
i
f
f
e
r
e
n
t
d
E
0
=
d
t
O
i
l
S
a
t
u
r
a
t
e
s
w
t
%
A
r
o
m
a
t
i
c
s
w
t
%
R
e
s
i
n
s
w
t
%
A
s
p
h
.
w
t
%
T
A
N
m
g
K
O
H
=
g
M
W
g
=
m
o
l
D
e
n
s
i
t
y
4
0

C
(
g
=
c
m
3
)
V
i
s
c
o
s
i
t
y
4
0
C
(
c
P
)
V
i
s
c
o
s
i
t
y
6
0

C
(
c
P
)
I
F
T
6
0

C
m
N
=
m
E

6
0

C
m
N
=
m
0
.
0
0
2
k
V
=
c
m
s

1
k
V
=
c
m
0
.
0
0
4
k
V
=
c
m
s

1
k
V
=
c
m
0
.
0
0
8
k
V
=
c
m
s

1
k
V
=
c
m
0
.
0
2
k
V
=
c
m
s

1
k
V
=
c
m
0
.
0
4
k
V
=
c
m
s

1
k
V
=
c
m
1
3
3
.
1
4
6
.
0
1
8
.
8
1
.
2
4
.
2
3
6
8
0
.
9
5
2
7
6
8
0
1
0
.
4
1
9
1
.
4
1
.
7
1
.
9
2
.
3
2
.
8
2
4
3
.
5
4
1
.
1
1
2
.
6
1
.
9
0
.
6
2
7
8
0
.
8
8
1
8
9
1
8
.
3
1
7
0
.
7
0
.
9
1
.
2
1
.
5
1
.
6
3
3
0
.
6
4
3
.
3
2
0
.
9
5
.
0
1
.
3
3
4
3
0
.
9
3
1
1
8
4
3
2
1
.
3
7
2
.
2
2
.
5
2
.
8
3
.
4
3
.
9
4
3
5
.
2
3
0
.
1
2
2
.
1
3
.
1
2
.
3
5
3
6
0
.
9
5
2
2
8
6
5
1
5
1
6
.
9
8
2
.
8
3
.
0
3
.
3
4
.
0
4
.
0
5
5
2
.
2
3
6
.
0
1
0
.
4
1
.
2
0
.
0
2
4
6
0
.
8
5
9
4
1
2
.
0
1
8
0
.
3
0
.
5
0
.
5
0
.
6
0
.
6
6
2
6
.
0
3
8
.
0
2
0
.
5
2
.
8
7
.
5
4
2
9
0
.
9
8
2
2
6
9
4
6
4
1
0
.
2
2
0
2
.
7
3
.
3
3
.
5
4
.
0
4
.
0
7
4
6
.
6
3
7
.
6
1
4
.
0
1
.
7
1
.
4
3
0
9
0
.
9
0
3
6
1
7
1
7
.
4
1
0
0
.
8
1
.
0
1
.
4
1
.
8
2
.
1
8
3
2
.
8
3
7
.
2
1
6
.
8
1
2
.
9
1
.
1
4
7
3
0
.
9
6
5
4
1
8
1
0
3
1
1
8
.
8
6
3
.
9
4
.
0
4
.
0
4
.
0
4
.
0
9
5
1
.
3
3
7
.
5
9
.
8
1
.
2
0
.
0
2
6
8
0
.
8
6
1
3
6
1
7
.
9
1
4
0
.
6
1
.
2
1
.
5
1
.
9
2
.
1
1
0
4
7
.
5
4
0
.
1
1
1
.
5
0
.
8
2
.
9
3
1
9
0
.
9
1
4
9
2
0
1
1
.
1
1
2
0
.
8
1
.
2
1
.
4
1
.
7
1
.
9
1
1
5
2
.
3
3
9
.
4
7
.
2
0
.
6
2
.
9
2
4
2
0
.
8
7
1
1
5
1
4
.
9
1
2
0
.
3
0
.
4
0
.
5
0
.
5
0
.
8
1
2
3
6
.
5
4
7
.
4
1
1
.
6
3
.
9
0
.
4
3
3
8
0
.
9
3
1
0
9
4
0
1
3
.
4
7
0
.
8
0
.
9
1
.
2
1
.
7
2
.
0
1
3
3
2
.
8
3
3
.
2
2
8
.
5
4
.
3
5
.
2
5
2
7
0
.
9
8
2
3
3
9
7
2
8
8
0
1
0
.
7
1
2
4
.
0
4
.
0
4
.
0
4
.
0
4
.
0
1
4
5
7
.
3
3
3
.
4
8
.
2
1
.
0
0
.
4
2
6
3
0
.
8
7
1
3
7
1
5
.
3
1
8
0
.
0
n
.
a
.
0
.
4
0
.
5
0
.
6
1
5
3
5
.
8
4
1
.
3
1
8
.
2
4
.
6
3
.
4
4
3
6
0
.
9
4
5
0
5
1
5
1
1
7
.
2
7
2
.
1
2
.
9
3
.
7
4
.
0
4
.
0
1
6
3
8
.
2
4
3
.
6
1
4
.
1
3
.
8
0
.
6
3
3
5
0
.
9
2
8
1
3
3
1
4
.
0
7
0
.
8
0
.
9
1
.
0
1
.
4
1
.
7
1
7
5
1
.
5
4
1
.
3
7
.
0
0
.
2
1
.
8
2
6
8
0
.
8
9
1
6
8
1
1
.
4
2
5
0
.
7
0
.
9
1
.
2
1
.
4
1
.
7
1
8
2
6
.
2
4
1
.
5
2
1
.
9
1
0
.
2
3
.
7
4
3
5
0
.
9
7
2
4
9
7
5
2
8
1
1
.
0
7
1
.
9
2
.
4
2
.
8
3
.
3
4
.
0
1
9
5
7
.
2
3
4
.
0
8
.
1
0
.
2
0
.
2
2
3
5
0
.
8
5
8
3
1
6
.
1
1
1
0
.
2
0
.
5
0
.
6
0
.
7
0
.
9
2
0
2
5
.
9
5
3
.
1
1
1
.
5
5
.
4
0
.
0
2
9
5
0
.
9
1
3
3
1
4
1
6
.
2
1
9
1
.
4
1
.
6
1
.
9
2
.
5
3
.
0
2
1
4
2
.
9
4
2
.
1
1
3
.
9
0
.
6
2
.
5
3
7
4
0
.
9
2
1
3
8
4
8
1
2
.
4
2
3
1
.
1
1
.
3
1
.
4
1
.
8
2
.
1
2
2
4
0
.
3
4
1
.
0
1
5
.
6
2
.
8
1
.
7
3
4
5
0
.
9
2
1
0
4
4
0
1
7
.
8
9
1
.
7
1
.
9
2
.
1
2
.
4
3
.
0
2
3
5
7
.
9
3
4
.
8
6
.
7
0
.
2
0
.
7
2
4
7
0
.
8
7
1
1
6
8
.
3
1
7
0
.
1
0
.
1
0
.
1
0
.
2
0
.
3
2
4
5
6
.
1
3
7
.
6
6
.
2
0
.
1
0
.
9
2
2
7
0
.
8
8
1
2
6
1
0
.
9
3
2
0
.
2
0
.
2
0
.
3
0
.
4
0
.
4
2
5
4
3
.
8
3
8
.
4
1
5
.
0
2
.
3
2
.
2
3
3
4
0
.
9
2
8
2
3
1
1
6
.
5
1
6
1
.
1
1
.
4
1
.
9
2
.
3
2
.
4
2
6
6
3
.
7
2
8
.
6
6
.
3
0
.
3
0
.
5
2
0
4
0
.
8
4
5
2
1
2
.
9
1
7
0
.
6
0
.
7
0
.
8
0
.
9
0
.
9
2
7
3
9
.
0
3
9
.
2
1
8
.
4
3
.
4
2
.
5
3
6
6
0
.
9
4
2
2
5
7
2
1
4
.
4
1
4
1
.
4
1
.
7
2
.
1
2
.
4
2
.
9
2
8
4
1
.
2
4
5
.
0
1
0
.
6
2
.
2
0
.
0
3
3
1
0
.
9
0
3
6
1
6
2
3
.
9
1
4
1
.
2
1
.
7
2
.
3
3
.
0
3
.
5
2
9
3
9
.
9
3
9
.
7
1
4
.
3
6
.
2
0
.
8
3
4
5
0
.
9
0
6
9
2
9
2
3
.
2
1
5
2
.
1
2
.
4
2
.
9
3
.
5
3
.
9
3
0
5
8
.
4
3
0
.
7
7
.
6
3
.
3
1
.
9
2
7
0
0
.
8
6
1
0
5
5
.
1
1
5
1
.
6
1
.
8
2
.
2
2
.
4
2
.
8
a
A
n
a
l
y
t
i
c
a
l
m
e
t
h
o
d
e
x
p
l
a
i
n
e
d
b
y
H
a
n
n
i
s
d
a
l
e
t
a
l
.
[
2
2
]
b
E
q
u
i
l
i
b
r
i
u
m
i
n
t
e
r
f
a
c
i
a
l
t
e
n
s
i
o
n
a
n
d
e
l
a
s
t
i
c
m
o
d
u
l
u
s
E

.
A
n
a
l
y
t
i
c
a
l
m
e
t
h
o
d
i
s
e
x
p
l
a
i
n
e
d
b
y
H
a
n
n
i
s
d
a
l
e
t
a
l
.
[
3
7
]
1445
D
o
w
n
l
o
a
d
e
d

b
y

[
L
'
O
r
e
a
l
]

a
t

1
2
:
1
0

1
0

M
a
r
c
h

2
0
1
4

Вам также может понравиться