You are on page 1of 51

An Article Submitted to

The Canadian Journal of Chemical


Engineering
Manuscript 1555

The Stability of Water-in-Crude and


Model Oil Emulsions
Peter K Kilpatrick∗


peter kilpatrick@ncsu.edu

Copyright 2005
c by the author. All rights reserved. No part of this publication may be re-
produced, stored in a retrieval system, or transmitted, in any form or by any means, electronic,
mechanical, photocopying, recording, or otherwise, without the prior written permission of the
publisher, bepress, which has been given certain exclusive rights by the author.
The Stability of Water-in-Crude and Model Oil
Emulsions
Peter K Kilpatrick

Abstract

The critical electric field (cef) technique has been utilized to measure the stabilities of a variety
of water-in-model oil and petroleum emulsions. The cef method allows for a fast, reproducible,
and quantitative gauge of emulsion stability. Here, we have used cef to measure the stability
of water-in-heptane-toluene-asphaltene emulsions and confirmed the importance of solvation of
asphaltenes and the state of asphaltene aggregation to emulsion stability. Emulsion stability in-
creased with the concentration of soluble asphaltenes near the point of precipitation. Droplet sizes
were measured with optical microscopy in order to calculate interfacial areas and film thicknesses.
It was found that film thickness increased with asphaltene concentration up to the solubility limit,
above which increased concentration had little effect, and cef increased with interfacial film thick-
ness up to a monolayer coverage of asphaltene aggregates, above which film thickness had a much
smaller effect. These findings were applied to a cef investigation of water-in-petroleum emulsions
to develop correlations of the stability of water-in-crude oil emulsions. A strong correlation (co-
efficient = 0.95) was found for cef with the product of asphaltene concentration and the difference
in hydrogen to carbon atomic ratios of the asphaltenes and petroleum solvent. The development
of a kinetic model and its fit to experimental data revealed the effects of asphaltene chemistry,
solvency, and resin concentration on the adsorption and consolidation of emulsion stabilizing in-
terfacial films. Keywords: Petroleum emulsions, asphaltene emulsions, critical electric field
Kilpatrick: The Stability of Water-in-Crude and Model Oil Emulsions 1

THE STABILITY OF
WATER-IN-CRUDE AND MODEL OIL EMULSIONS

Andrew P. Sullivan, Nael N. Zaki, Johan Sjöblom#,


and Peter K. Kilpatrick*

Department of Chemical and Biomolecular Engineering


North Carolina State University
Raleigh, North Carolina 27695-7905
(919) 515-7121; (919) 515-3465 (facsimile)
peter-k@eos.ncsu.edu

#
Ugelstad Laboratory
Department of Chemical Engineering
Norwegian University of Science and Technology
N-7491 Trondheim, Norway

Submitted to The Canadian Journal of Chemical Engineering

As part of the Special Issue Honoring Professor Jacob H. Masliyah

June 2007

* Author to whom correspondence should be addressed

Produce by The Berkely Electroni Pres, 205


2

Abstract

The critical electric field (cef) technique has been utilized to measure the
stabilities of a variety of water-in-model oil and petroleum emulsions. The cef method
allows for a fast, reproducible, and quantitative gauge of emulsion stability. Here, we
have used cef to measure the stability of water-in-heptane-toluene-asphaltene emulsions
and confirmed the importance of solvation of asphaltenes and the state of asphaltene
aggregation to emulsion stability. Emulsion stability increased with the concentration
of soluble asphaltenes near the point of precipitation. Droplet sizes were measured
with optical microscopy in order to calculate interfacial areas and film thicknesses. It
was found that film thickness increased with asphaltene concentration up to the
solubility limit, above which increased concentration had little effect, and cef increased
with interfacial film thickness up to a monolayer coverage of asphaltene aggregates,
above which film thickness had a much smaller effect. These findings were applied to
a cef investigation of water-in-petroleum emulsions to develop correlations of the
stability of water-in-crude oil emulsions. A strong correlation (coefficient = 0.95) was
found for cef with the product of asphaltene concentration and the difference in
hydrogen to carbon atomic ratios of the asphaltenes and petroleum solvent. The
development of a kinetic model and its fit to experimental data revealed the effects of
asphaltene chemistry, solvency, and resin concentration on the adsorption and
consolidation of emulsion stabilizing interfacial films.

Keywords: Petroleum emulsions, asphaltene emulsions, critical electric field

2
Kilpatrick: The Stability of Water-in-Crude and Model Oil Emulsions 3

1. Introduction

The existence of emulsions in both the production and refining of crude oil is
one of the most persistent problems facing the petroleum industry. At the production
end, emulsions are generated when water is coproduced with petroleum or pumped into
the well to aid in petroleum recovery and this mixture passes through the wellhead, pipe
bends, and choke valves. At the refinery, water is added to generate a large oil-water
interfacial area to aid in the extraction of salts from the crude oil. The emulsions
produced from these processes do not easily resolve into neat crude and water phases
1-5
and a certain volume of the emulsion remains. The emulsified water can corrode
refinery equipment such as distillation columns and dissolved salts in the water can
5
poison catalysts. Some emulsions are very viscous and will foul machinery if allowed
to continue through the refining processes. To minimize these problems, emulsions
undergo separation procedures such as electrostatic coalescence and demulsifier
addition. There is a great deal of ongoing research attempting to correlate crude oil and
6-18
demulsifier properties with the effectiveness of these separation methods.

1.1. Surface-Active Components in Petroleum

The mechanism of water-in-crude oil emulsion stabilization is not fully


understood. The primary means, however, by which water droplets are stabilized
appears to be the formation of a viscoelastic, mechanically strong film at the droplet
19-25
interface composed of asphaltenes. Asphaltenes are the portion of petroleum
insoluble in an n-alkane solvent (usually n-pentane or n-heptane). Asphaltenes are
polydisperse with respect to molecular weight and chemical functionality. The
common structural theme among asphaltenes is a planar, polyaromatic, fused ring core
imbedded with polar functionality, surrounded by aliphatic side chains and naphthenic
26-32
rings (Figure 1a). In petroleum, asphaltene molecules aggregate to minimize the
interactions between their polar cores and the non-polar solvent and to increase the
27,33,34
extent of -bond overlap. Asphaltenes are surface-active, having hydrophobic
and hydrophilic portions, however, unlike typical surfactants, asphaltenes and their

3
Produce by The Berkely Electroni Pres, 205
4

aggregates likely orient parallel to the interface, exposing their cores to the water phase.
Scanning tunneling microscopy has been utilized to image the surface of pyrolytic
35
graphite with adsorbed asphaltenes. The surface was highly ordered and flat with
periodic features extending from the surface ca. 2 Å. The results indicated asphaltene
sheets oriented parallel to the surface with the terminal carbons of the aliphatic side-
chains extending from the surface. The orientation of asphaltenes upon exposure to
graphite may be similar to that with water interfaces, both surfaces preferentially
interact with asphaltenic cores.
Resins, another class of surface-active molecules in petroleum, have their
hydrophobic and hydrophilic portions on opposite ends, encouraging interfacial
adsorption. Resins solvate asphaltenic aggregates by adsorbing to the faces of these
34,36,37
aggregates, a process sometimes referred to as peptization (Figure 1b) The
interaction of asphaltenes and resins and the subsequent adsorption to the water-oil
interface forms a rigid, viscoelastic, cross-linked network preventing droplet
19
coalescence (Figure 1c). Mackay and coworkers observed this structure when they
formed a water droplet in an asphaltene-containing oil phase, allowed time for
adsorption to the interface, and withdrew the water. This process yielded a sac-like
structure that folded as the water was removed but did not rupture.
Emulsions produced via asphaltene-resin aggregation and adsorption to the
interface are a major concern to the petroleum industry. As oil is produced, the lighter
material with lower asphaltene and water content is encountered first. As wells near
completion, the light end supplies are exhausted, and the heavier, higher asphaltene
content petroleum is produced. As heavier crudes are produced, the industry will have
to develop ways to better handle and understand asphaltene physical chemistry and the
emulsion problem.

1.2. Methods of Measuring Emulsion Stability

There have been many attempts at quantifying and correlating emulsion


6,11,12,38-49
stability. So-called bottle tests and droplet size observations are common. In
bottle tests, an emulsion is created and the volume of the disperse phase resolved over

4
Kilpatrick: The Stability of Water-in-Crude and Model Oil Emulsions 5

time is observed. To produce emulsions that will show resolution in a relatively short
period of time, the level of mixing energy must be low, or the composition of the oil
system must be one that produces weak emulsions. Both strong and weak emulsions
resulting from a range of mixing energies and crude compositions are encountered in
actual production and refinery conditions. Another possibility is to produce emulsions
with a high mixing energy and propensity to emulsify and measure the water resolved
with the application of a centrifugal field. This will produce a range of resolved water
volumes, however this technique is very dependent upon protocol and only provides
relative stabilities. The measure of the volume of water resolved due to coalescence of
water droplets in a high centrifugal field may not be a true measure of real emulsion
50
strength. In refinery situations, centrifugation on a large scale for all emulsions is
very expensive and not widely performed. Therefore, any predictive correlation with
this technique as a basis is of limited value.
Droplet size is commonly measured by means of optical microscopy or light
51,52
scattering techniques. Optical microscopic observations of emulsion samples are
difficult to perform without disturbing the emulsion system. In order to generate an
accurate count and droplet size distribution, the emulsion droplets must be confined
between a glass slide and cover slip to provide an unobstructed view of one droplet
layer. This distorts droplet size and may result in some droplet coalescence, producing
an inaccurate droplet count and size distribution.
50,53-56
The use of light scattering techniques is troublesome as well. Petroleum
is opaque so a very thin sample of emulsion must be observed. The observable path
length may be very small for high asphaltene content systems, complicating sample
preparation and observation. The conversion of light scattering data into droplet size
information requires the assumption of a droplet size distribution. Lastly, light
scattering techniques can not distinguish between coagulated and individual water
droplets. In stable petroleum emulsion systems, many droplet collisions do not result in
coalescence and coagulated droplets are common. With all of these difficulties, light
scattering has somewhat limited value for this application.
The primary mechanism of emulsion stabilization is the formation and stability
of the interfacial film discussed above. Therefore, a useful, predictive correlation

5
Produce by The Berkely Electroni Pres, 205
6

would be one that probes this property directly. In this study, emulsion stability was
quantified with the determination of the electric field strength required for emulsion
breakdown in w/o emulsions. The technique is based on work by Sjöblom and
57-61
coworkers with modifications in sample cell design, emulsion sampling procedures,
and electronic components. To measure stability, an emulsion sample is placed
between two electrodes and the voltage is steadily increased (Figure 2). At low
voltages, a low level of current is observed due to conduction through the continuous
oil phase. As the voltage is increased, water droplets begin to coagulate in small chains
parallel to the electric field due to polarization of the water droplets from the movement
62,63
of electrolytes in the water phase. As the voltage is increased further, the small
chains form bridges spanning the entire gap between the electrodes. Finally, at a
certain electric field strength, the electromotive force exerted on the electrolyte ions is
sufficient to rupture the interfaces separating the water droplets. This phenomenon is
observed as a large change in slope of the current versus voltage curve. The electric
field (in kV/cm) at which this occurs is termed the critical electric field (cef).
Cef is a direct measure of the strength of the asphaltene–resin interfacial film
and is simple to perform. Emulsion droplets encounter similar forces at the refinery in
ac electrostatic coalescers. In these units, droplets coalesce due to increased collision
64,65
and thinning of interfaces as droplets are deformed in an alternating electric field.
In the cef technique, a dc power supply is used and coalescence does not occur by the
same mechanism, but the important feature, thinning or disruption of the interface in
the presence of an electric field, is identical. Therefore, no artificial artifacts of the
experimental procedure are introduced in the measure of emulsion stability and it can
easily be adapted to field work and testing. The technique produces a unique number
for a given emulsion and is independent of the choice of experimental conditions which
do not directly affect emulsion stability (gap width, sample size, etc.), unlike other
techniques which are dependent on sample handling and testing procedures. Cef only
depends on variables such as temperature, water volume fraction, droplet size
distribution, asphaltene content, etc. which control emulsion stability. The measurable
range of emulsion stability is large (0.01-5 kV/cm) and the test is rapid and
reproducible. However, the cef technique does not provide an answer as to the extent

6
Kilpatrick: The Stability of Water-in-Crude and Model Oil Emulsions 7

of the water phase that separates upon demulsification. Neither do we get any
information about the stability of the residual emulsion droplets remaining after field-
induced coalescence.
In this study we have used the cef technique to develop a water-in-petroleum
emulsion stability correlation with properties of the petroleum using information
gathered from an investigation of cef in model oil systems. These tests have been
performed at two different water/oil ratios producing different results. The
development of this correlation has revealed important insight into the driving
mechanisms for asphaltene aggregation and emulsion stabilization. Observations of cef
versus time for many petroleum emulsions have led to the development of a simple
kinetic model for the build-up of the interfacial film due to asphaltene adsorption that
has revealed information on the effects of asphaltene chemistry and solvency. The
similarities of petroleum results with those of model oils show that the observed effects
with petroleum truly are due to asphaltenic films.

2. Experimental

2.1. Materials

A variety of crude oils were used for this study. Properties important for
emulsion stability are listed in Table 1 for the crude oils investigated as well as others
present in our laboratory. One important set of properties is the asphaltene and resin
content. For the crude oils studied, these values range from 0.79% to 14.8% (w/w) for
asphaltenes and 3.24 to 20.5% (w/w) for resins, yielding a relatively narrow range of
resin to asphaltene ratios (R/A = 0.90-4.25, except AL, R/A = 6.17). The asphaltene
content was measured by precipitation in n-heptane and the resin content was obtained
49
by sequential elution chromatography of the deasphalted crude adsorbed on silica gel.
Another important set of properties for the crude oils is the hydrogen to carbon
atomic ratio (H/C). This number provides a quantitative gauge of the aromatic vs.
aliphatic nature of a chemical species (e.g. H/Cheptane=2.29, H/Ctoluene = 1.14). The

7
Produce by The Berkely Electroni Pres, 205
8

ratios reported in the Table correspond to the whole crude oil, the asphaltene fraction,
the resin fraction, the crude after removing the asphaltenes (DeA crude), and the
difference in H/C of the DeA crude and asphaltenes ( H/C). The hydrogen and carbon
contents were measured by elemental analysis performed by Galbraith Laboratories.
The H/C ratios of the DeA crudes were calculated from hydrogen and carbon contents
of the whole crudes, asphaltenes, and the concentration of asphaltenes in the whole
crude. Elemental analyses of asphaltenes were performed with a Perkin Elmer 2400
Series II CHNS/O Analyzer.
Other properties reported include density, which varies from 0.833 to 1.01
g/mL, and viscosity, which ranges from 3.14 cP to 2310 cP at 100 ˚F (except for SCS, a
waxy crude with an unmeasurable viscosity below its pour point). These properties
control the shear field in the oil sample during emulsification. Droplets formed during
emulsification immediately begin to coalesce as they collide with other droplets if not
stabilized by an asphaltenic film. When emulsification is completed, new droplets are
no longer formed but coalescence continues between unstable droplets. Asphaltenic
films build up over time and resist coalescence to a greater extent at longer times. The
balance between the rates of asphaltene adsorption and droplet-droplet collisions
determines the ultimate droplet size distribution and stability of the emulsion. As a
result, factors that govern asphaltene and water droplet mobilities in the oil phase may
be important to consider for the development of an emulsion stability correlation. For
the crude oils in this study, density was measured pycnometrically, the viscosity was
measured with a Rheometrics couette rheometer, and the kinematic viscosity was
calculated from these two numbers.
Deionized water with 1 % (w/w) NaCl added was used for creating the
emulsions. The salt was necessary for the measurement of critical electric field. After
NaCl addition, the pH of the water was adjusted to 6 with dilute HCl and NaOH.
Model oils used in this investigation consisted of n-heptane and toluene with
asphaltenes fractionated from Arab Heavy (AH) [also known as Safaniya] crude oils.
All solvents used were HPLC grade and supplied by Fisher Scientific.

2.2. Emulsion Preparation

8
Kilpatrick: The Stability of Water-in-Crude and Model Oil Emulsions 9

Prior to emulsification, petroleum was loaded into a metal cylinder that was
sealed and placed in a 100˚C oven for two hours in order to erase the thermal history
and ensure the complete melting of wax crystals. The cylinder was then placed in a
60˚C water bath for 30 minutes, after which a sample was transferred to a polyethylene
jar with the appropriate amount of water (also at 60˚C). Emulsions were prepared at
two different oil/water ratios: 4/6 and 7/3 (v/v). The crude oil and water were mixed
for 5 minutes (3 minutes at the level of the original bulk oil/water interface and 2
minutes at the bottom of the jar) at 15,000 rpm with a Virtis VirtiShear Cyclone IQ
Homogenizer, using a 6 mm diameter internal shaft rotor/stator assembly with a gap
width of 0.127 mm. The mixing was performed in a water bath to maintain the system
at 60 ˚C. After emulsification, the sample was placed in a 60 ˚C oven until stability
testing was performed.
For model oil experiments, AH asphaltenes were dissolved in toluene for 2
hours. After this period, heptane was added to adjust the solvent to the appropriate
degree of aromaticity. The model oils (referred to as heptol) varied from 10-60 %
toluene (v/v) in heptane. After 2 hours of mixing, the appropriate amount of water was
added and the mixture was emulsified for 3 minutes (2 minutes at the bulk oil/water
interface and 1 minute at the bottom of the jar). The temperature was not controlled for
the model oil experiments because no wax was present as in the crude oil.

2.3. Critical Electric Field Measurement

The stability of the water-in-crude oil emulsions was gauged with critical
electric field (cef) measurements. To measure cef, an emulsion sample was placed in
the sample cell (in a 60 ˚C oven) consisting of two, 1.0 cm diameter, gold plated,
copper electrodes, separated with Mylar spacers and held in an aluminum casing.
Figure 3 shows a side view of the cell. The cell was designed so the gap width could be
varied but for all of the experiments it was 0.250 mm. Two holes were drilled through
the top for sample introduction. A syringe was used to withdraw a sample from the
middle of the emulsion and inject it through one of the holes in the cell. The cell was
connected to a HP6634B power supply (0-100 V dc source), controlled by a PC through

9
Produce by The Berkely Electroni Pres, 205
10

the use of a HP82350A interface card. Using this card, the power supply was
controlled with a Visual Basic program.
After loading the sample in the cell, the voltage between the electrodes was
increased in increments of 0.25 V every 5 seconds and the current was measured 2
seconds after every step change (to avoid current spikes). All of the emulsions for
which we have reported cef showed no water resolved upon visual inspection after a 24
hour period. The presence of resolved water would have indicated an emulsion
containing large, unstable droplets. All of the emulsions tested were relatively stable
and had droplet diameters in the range of 0.5-20 µm. Droplet size appears to be an
important parameter in the stability of emulsions gauged by cef as we will show from
the model emulsion results. Because of this sensitivity to droplet size, the emulsions
were always sampled from the middle of the emulsion. Microscopic observations were
performed in parallel with cef measurements to verify valid sampling procedures. All
of the crude oil systems appeared to have droplet size ranges that were close enough to
ensure valid comparisons from sample to sample. Model oil emulsion studies were
conducted in the same manner except all steps were performed at room temperature.

10
Kilpatrick: The Stability of Water-in-Crude and Model Oil Emulsions 11

2.4. Solubility Tests

The solubility of asphaltenes in various solvents was tested by preparing a 1.5 mL


model oil sample with 1 % asphaltenes (w/w) and injecting it through a glass
microfiber filter disk attached to the end of a syringe (1.6 µm pore size). After the
sample was filtered through the disk, a 1.5 mL aliquot of the model oil solvent was
pushed through to rinse any soluble material that may have adsorbed to the microfibers.
The rinse step was performed quickly in order to prevent solubilization of any
previously precipitated material. Both the filtrate and rinse were collected in a small
vial. The asphaltenes in the syringe tip and vial cap that were not dissolved in the
solvent were rinsed with methylene chloride into the original vial and labeled as
precipitate. The filtrate and precipitate vials were dried in a nitrogen flushed vacuum
oven at 70 ˚C for two days and weighed.

3. Results and Discussion

3.1. Asphaltene-stabilized model oil emulsions

3.1.1. 60 % water
Initial model experiments were performed with organic solutions of varying
heptane and toluene content, into which asphaltenes from Arab Heavy crude oil were
dissolved. These solutions of varying toluene % and AH asphaltene content were
emulsified with water such that the water content was 60% (w/o) and then the cef was
measured after ageing for 24 hours. The results, along with the solubility of AH
asphaltenes, are shown in Figure 4. The cef is observed to achieve local maxima with 2
and 3 % asphaltene solutions at a toluene concentration of ca. 40%, very close to the
solubility limit. This maximum in stability shifts to lower toluene concentrations at an
asphaltene concentration of 5 %. With the lower (2-3%) asphaltene concentrations, the
maximum stability occurs at the limit of solubility, the point at which the asphaltenes
are most surface-active and labile due to the fine aggregate size. Beyond the solubility
limit at these asphaltene concentrations, sufficient inventory of asphaltenes precipitate,

11
Produce by The Berkely Electroni Pres, 205
12

producing flocs which are much more weakly surface-active and insufficiently labile to
self-assemble at the interface producing a strong interfacial film. At higher asphaltene
concentrations (5-7%), the stability achieves a local maximum slightly beyond the
solubility limit because the greater inventory of material enables the droplets to be
better coated with asphaltenic film, despite the fact that some material precipitates
beyond the solubility limit. This phenomenon is observed with high water content
emulsions (e.g. 60%) in which there is a dearth of asphaltenic material until sufficiently
high asphaltene concentrations are reached (as we will show below).

3.1.2. 30 % water
Cef results for model emulsions produced with AH asphaltene solutions in
heptol and 30 % water are shown in Figure 5. Cef values for these emulsions were
higher than those with 60 % water. The total interfacial area was about the same as that
found with 60 % water, however the average droplet size was smaller. Larger droplets
may facilitate droplet chaining between electrodes or droplet coalescence in the electric
field. Cef values are similar for 0.5-3 % AH asphaltenes at toluene concentrations
above the solubility point, but are much lower at 0.25 % AH asphaltenes. This suggests
that there is some critical concentration of asphaltenes above which interfacial
thickness is large enough to provide a strong barrier for droplet coalescence and
increased asphaltene adsorption above this has little effect on emulsion stability. As for
the 60 % water emulsions, we see the maximal cef occurs at lower toluene
concentrations as the asphaltene concentration is increased. The peak locations are
about the same as those observed for AH asphaltenes with 60 % water but in this case
more solvent compositions were probed and more detailed shifts can be observed.
57
Sjöblom and coworkers obtained slightly higher cef values with similar types
of experiments. For model emulsions composed of 2 % asphaltenes in decane-toluene
mixtures, they found cef values ranging from 2.9 kV/cm at 20 % toluene to 0.55 kV/cm
at 80 % toluene. Emulsions were unstable at 100 % toluene. Details were not provided
of the asphaltene type or water content of the emulsions, but the magnitude of the cef’s
and trend with solvent aromaticity they observed support the results we have obtained
here.

12
Kilpatrick: The Stability of Water-in-Crude and Model Oil Emulsions 13

3.1.3. Interfacial film thicknesses


Droplet sizes generally decreased for all solvency conditions as asphaltene
concentration increased (Figure 6). Generally speaking, one would expect increasing
the toluene concentration to decrease the interfacial tension and thus reduce the droplet
size of water-in-oil emulsions, all other things being equal. However, as toluene
concentration is increased, the surface activity of the asphaltenes decreases and this has
a greater impact than interfacial tension. All droplet size data on Figure 6 are for the
soluble regime of AH asphaltenes. As expected, the droplet sizes decreased as
asphaltene concentrations were increased for all of the solvents shown. For cef testing
and droplet size observations, emulsions were sampled from the middle so the results
may not be representative of the overall emulsion. For all other asphaltene
concentrations, the droplet sizes decreased with concentration up until a critical
concentration.
The total water-oil interfacial area was calculated assuming a monodisperse
system:

(3ml)(1cm 3 /ml)(1m/100cm) 3 9x10 6 m 3


A w/o = N d A d = (4 R d2 ) = (1)
4 Rd
R 3d
3
where Nd is the number of water droplets, Ad is the surface area of each droplet, Rd is
the droplet radius, and Aw/o is the total water-oil interfacial area. Film masses were
calculated assuming 10 % adsorption of asphaltenes based on interfacial film studies
67
performed in our lab at conditions of maximum interfacial activity. This number is
probably dependent on the asphaltene solubility conditions, but the range of solvent
compositions was small enough that this is a good approximation.
Figure 7 displays the results of interfacial mass/area calculations for 45-55 %
toluene with AH asphaltenes (30 % water) and 50 % toluene with HO asphaltenes (60
% water). Interfacial masses increased with asphaltene concentration. The slopes of
the lines are similar for all cases for asphaltenes in the soluble regime. For HO
asphaltenes in 50 % toluene, cef did not increase past 2 % asphaltenes due to solubility
limitations. All of the trends with AH asphaltenes are linear up to 3 %. These results

13
Produce by The Berkely Electroni Pres, 205
14

are consistent with the solvency of the asphaltenes observed in previous experiments
(Figures 4 and 5). Figure 8 displays cef versus interfacial mass/area for all of the
solvency conditions in Figure 7. Cef increased with interfacial mass/area up to about
1.50 mg/m2, above which the dependence was much less. To calculate a rough estimate
of the extent of interfacial coverage in these emulsions, we assumed an asphaltene
molecular weight of 1,000 g/mole and asphaltene molecular dimensions of 0.5 nm thick
with a diameter of 2.0 nm. The molecular weight of asphaltenes has been an
68-76
extensively researched subject. Many techniques including, vapor pressure
osmometry, gel permeation chromatography, viscometry, and mass spectrometry have
been utilized to measure molecular weights with widely varying results. Numbers
from 900 to 18,000 g/mole, have been recorded, with the lowest values found with the
best solvents. An asphaltene molecular weight of 1,000 g/mole was used in our
analyses because it is at the low end of observed weights and probably corresponds to
individual asphaltene molecules. The asphaltene molecular dimensions used were
27
taken from x-ray diffraction results of Yen. We assumed an imperfect stacking of
asphaltenes so the asphaltene stack diameter (25 Å) is 25 % larger than that of the
individual molecule (20 Å). With these assumptions, the number of molecules
corresponding to the interfacial thickness was calculated:

0.00150 g / m 2
= 1.50 x10 6 mole / m 2 (2)
1000 g / mole
6 2 23 17 2 (3)
(1.50x10 mole/ m )(6.022x10 molec / mole) = 9.03x10 molec / m

(9.03 x1017 molec / m 2 )[ (1.25 x10 9 m) 2 / stack ] = 4.4molec / stack (4)

This number seems reasonable. Asphaltene aggregates consist of about 5 molecules in


27
the best solvent conditions, so the critical interfacial concentration appears to be
roughly a monolayer of asphaltene aggregates. The calculated interfacial thicknesses in
terms of number of molecules are displayed in Figure 8.

3.1.4. Kinetic Model for Interfacial Film Formation

14
Kilpatrick: The Stability of Water-in-Crude and Model Oil Emulsions 15

All of the cef measurements reported up until now were performed 24 hours
after emulsification. Greater differences among the various asphaltene concentrations
were shown at shorter times in which the asphaltenic film did not have time to reach the
critical thickness. Figure 9 shows cef results for model emulsions at various times from
approximately 15 seconds to 24 hours after emulsification. The cef for emulsions with
1 % asphaltenes at 40 % toluene changed very rapidly during the first half hour and
slowly at longer times. The cef was 1.04 kV/cm after 30 minutes and reached 1.24
kV/cm after 24 hours. For 0.5 % asphaltenes and 40 % toluene, the cef value did not
change in the first 30 minutes, but then the change in cef with time was very similar to
that observed for 1 % asphaltenes. At short times, the concentration of adsorbed
asphaltenes was too low to stabilize droplets and asphaltenes adsorbed to the interface
but did not appreciably affect emulsion stability. After 30 minutes, the coverage
reached a high enough concentration that increased adsorption led to increased stability
at the same rate as the 1 % asphaltene, 40 % toluene condition. The results obtained for
1 % asphaltenes and 50 % toluene were very different from either of the 40 % toluene
results. In this case the cef increase was much more gradual over the whole 24 hour
period. The effect of resins in model emulsions with 1 % AH asphaltenes and 50 %
toluene is also shown in Figure 9. The addition of 1 or 2 % AH resins resulted in weak
emulsions for all times up to 24 hours. These results will be discussed in terms of a
kinetic model for interfacial film formation in the next section.
The results for the model systems are similar in nature to those obtained by
77
Sjöblom and coworkers. Using a 50:50 mixture of “condensate F” and water with
various combinations of asphaltenes and resins (0.9-5 % A, 1-10 % R), they observed
cef over a period of one week. Cef values were found to generally increase over the
whole period, with the majority of the increase occurring in the first day. The
maximum electric field measured was 2.00 kV/cm, and is comparable to our findings.
The model for asphaltene adsorption and the stabilization of water-oil interfaces
consists of two processes. First, the adsorption of asphaltene aggregates (A) to the
interface and second, the consolidation of the adsorbed asphaltenes into a rigid, cross-
linked, interfacial structure78. The adsorption mechanism is as follows:
diffusion ka
Bulk A(c A ) Near Interface A(c As )
kc
(5)
Adsorbed A( 1 ) Consolidated A( 2)

15
Produce by The Berkely Electroni Pres, 205
16

The terms in parentheses are the concentrations of asphaltenes at each condition. Bulk
asphaltenes undergo Fickian diffusion to the interface where they immediately adsorb.
The adsorbed asphaltenes then consolidate to form an interfacial film. Film
consolidation involves the combination of an unconsolidated asphaltene molecule with
the consolidated film. The adsorption and consolidation steps, when asphaltenes are
adsorbing, are assumed to be irreversible and have rate constants of ka and kc
respectively. This process is modeled with the following set of differential equations:
d 1
= k a c As ( max 1 2 ) kc 1 2 (6)
dt
d 2
= kc 1 2 (7)
dt
1/ 2
Dt
wherec As = 2c bulk
A (8)

kT
and D = (9)
6 R

max is the maximum concentration of asphaltenes that can adsorb to the interface, cAbulk
is the concentration of asphaltenes in the bulk oil phase, D is the diffusivity of
asphaltenes in the oil, k is Boltzmann’s constant, is the viscosity of the oil, and R is
the radius of asphaltene aggregates. The equation for concentration of asphaltenes near
the interface was derived from Fick’s law for diffusion assuming the bulk concentration
remained constant. For the calculation of asphaltene diffusivity, the aggregate diameter
was assumed to be 20 Å and is based on observations of Yen in model oil systems. As
asphaltenes adsorb, the driving force for further adsorption is reduced due to
asphaltene-asphaltene repulsion and shielding of the water phase. This is accounted for
with the saturation term, max. Under conditions of identical dispersed water content,
we assume cef is proportional to the total amount of adsorbed asphaltenes. The data
show that cef increases very rapidly at first followed by a more gradual increase with
time, suggesting that adsorbed asphaltenes immediately provide emulsion stability and
consolidation contributes to a lesser extent over a long time period. To model this
behavior we set cef equal to a linear combination of 1 and 2:
CEF = 1 1 + 2 2 (10)

16
Kilpatrick: The Stability of Water-in-Crude and Model Oil Emulsions 17

1 + 2 =1 (11)
The cef value at the first point (3 minutes) was fit by assuming 1 and 2 were equal at
that time. The values of 1 and 2 were then calculated:
cef(3 minutes) = 1 1 (3 minutes) + 2 2 (3 minutes) (12)
cef(3 minutes)
1 (3minutes ) = = 2 (3minutes ) (13)
1 + 2

1 (3 minutes) = 2 (3 minutes) = cef(3 minutes) (14)


To calculate max, the interface was assumed to be saturated with consolidated
asphaltenes at 24 hours.
cef(24 hours) = 2 2 (24 hours) = 2 max (15)

cef(24 hours)
max = (16)
2

The above set of equations could not be solved analytically. Runge-Kutta


numerical methods were used to generate a series of parametric curves in which the
parameters were systematically varied to find the best fit with the data. The best ka/kc
combinations for several values of 1 and 2 were obtained by minimizing the sum of
the squares of the differences between the experimentally measured cef and model
prediction values for all times. The optimized fit parameters are listed in Table 2. For
all model and crude oils, values of 0.2 and 0.8 for 1 and 2, respectively, provided a
good fit to the data.
The kinetics of three different model systems were investigated: 0.5 and 1.0 %
AH asphaltenes in 40 % toluene, and 1.0 % AH asphaltenes in 50 % toluene. The
model was unable to predict a high enough rate of adsorption to account for the very
rapid cef increase for 0.5 and 1.0 % AH asphaltenes in 40 % toluene. At these
conditions, the driving force for asphaltene adsorption is very large due to the large
chemical mismatch and high concentration of precipitated aggregates. The lyophobic
forces associated with the asphaltene aggregates in these conditions may not be
accounted for in the model. The fit of the model to the 1.0 % AH asphaltene in 50 %
toluene results provided a ka of 0.8 cm3min-1g-1 and a kc of 0.02 min-1. The rate constants
could not be derived for the model oils with resins. The near constant low cef
prohibited an accurate fit of the model. It is expected that resins would decrease the

17
Produce by The Berkely Electroni Pres, 205
18

driving force for asphaltene adsorption to the interface. The role of resins will be
explored more with petroleum emulsions.
From model emulsion studies we have confirmed the importance of the cef
technique through a correlation of interfacial film thickness with the magnitude of the
cef and the construction of a kinetic model for interfacial film formation that fits the
data. We have also learned the importance of asphaltene concentration and solvency as
gauged by • H/C on emulsion stability. We will now use these findings to develop a
correlation for petroleum emulsion stability and to discover the parameters that govern
interfacial film development.

3.2. Petroleum Emulsions


3.2.1. Determination of Petroleum Emulsion Stability Correlation
Cef was measured for emulsions prepared with 30 % water (v/v) for 12 different
petroleums after ageing for 24 hours. All 12 petroleums produced emulsions that were
at least stable to gravity over the 24 hour ageing period. The decreased water content
led to greater distances between emulsified water droplets immediately after
homogenization, resulting in lower droplet-droplet collision frequencies, allowing
longer times for interfacial film development. Based on the model oil emulsion results,
we know asphaltene solvency plays a large role in emulsion stability. Additional
correlations in which resin content appeared with a negative exponent and the
difference between the H/C ratio of the asphaltenes and petroleum solvent appeared
with a positive exponent were attempted. The dependence of cef on the product of
asphaltene concentration and H/C for petroleum emulsions, prepared with 30 % water
(v/v), after 24 hours of ageing, is displayed in Figure 10. A correlation coefficient of
0.88 was obtained when South China Sea crude was included. Much of the deviation
from linearity is attributable to the point corresponding to South China Sea (SCS). This
particular crude contains over 32% wax. The experiments were performed at 60 ˚C, but
a small percentage of the wax in SCS may be precipitated at this condition.
Alternatively, the extent of asphaltene solvation, as measured by H/C, in SCS may
not be comparable to that found in the other crudes. The effect of the long paraffinic

18
Kilpatrick: The Stability of Water-in-Crude and Model Oil Emulsions 19

wax molecules in SCS on asphaltene solubility differs significantly from the effect of a
shorter, normal alkane which is not accounted for with the H/C ratio. For example
H/C’s of n-decane and n-triacontane (C30) are very similar (2.20 vs. 2.07, respectively)
despite the fact that they differ by twenty carbons and result in differences in asphaltene
solubility (longer alkane chains solubilize asphaltenes better than shorter ones). In
addition, the branched nature of the long chain waxes found in SCS lowers the H/C
relative to straight chain molecules with the same carbon number without significantly
changing its effect on asphaltene aggregation. Therefore, SCS was removed from the
correlation and the coefficient improved from 0.88 to 0.95. The correlation does not
pass through the origin, suggesting the existence of a critical asphaltene concentration
below which, emulsions are not stabilized, as seen for model emulsions. To calculate a
rough estimate of the extent of interfacial coverage in these emulsions, we assume, as
before, an asphaltene molecular weight of 1,000 g/mole, 20 Å diameter asphaltene
stacks, and a crude oil density of 0.9 g/mL. The total water-oil interfacial area was
calculated as before. Based on this area, the total droplet coverage by asphaltenes was
calculated:
m A = (7mLoil )(0.9 g / mL)(%wt A / 100)(0.1 g A adsorbed/g A) = 0.0063(%wt A) g (17)
[0.0063(% wt A) g]
(6.022x10 23 molec / mole)
1000 g / mole
N AA = = 7.6x1017(% wt A) agg (18)
5 molec / agg
S ads = S AA N AA = (1.0 x10 9 m) 2 [7.6 x1017 (%wt A)] = 2.4(%wt A) m 2 (19)

2.4(%wt A)
= 6 3
= 2.7 x10 5 (%wt A)( Rd ) m -1 (20)
9 x10 m
Rd

where mA is the mass of asphaltenes. NAA and SAA are the number and cross-sectional
area of asphaltene aggregates, assuming columnar stacking of asphaltene molecules in
27
groups of five. Sads is the total surface area covered by asphaltenes and is the
fraction of interfacial area covered. The crude oils used had a range of asphaltene
concentrations from 0.79 to 14.8 % corresponding to monolayer coverage of droplets
with diameters of 9 to 0.5 µm. Size distributions were troublesome to perform for
petroleum emulsions and prone to significant errors because the emulsions were

19
Produce by The Berkely Electroni Pres, 205
20

opaque. Droplet counts were not performed as for the model emulsions but qualitative
observations were made. All of the emulsions appeared to have average diameters a
little greater but close to these values. The smallest droplets observed in our
experiments were about a micron in diameter, implying that this was the predominant
size immediately after emulsification before any coalescence occurred. At this point,
coalescence was fast and droplet diameters increased rapidly, because the extent of
interfacial coverage was low. As droplet diameters increased, the interfacial area
decreased, providing a higher % coverage. At long times the coalescence rate became
very small as multi-layered interfacial films covered the entire interfacial area. Given
the fact that asphaltene interfacial adsorption is not instantaneous, it is not surprising
that the actual droplet size distribution is slightly larger than that predicted from
monolayer coverage. All of the systems reported here are within the range of what we
would term the “minimum drop size distribution” possible and are controlled by the
extent of interfacial film formation rather than the extent of mixing.
The above correlation provides a remarkably good fit to the data without
involving many petroleum characteristics considering the large variety of samples
studied and possible correlating parameters. Resins play a large role in the
solubilization of asphaltenes and would be expected to have a large impact on emulsion
stability. All of the petroleums included in the correlation have R/A values within a
relatively narrow range (0.92-4.25, except for Alba = 6.17) and the extent of resin
solvation of asphaltene aggregates may be fairly uniform. For petroleum with very
different R/A values, emulsion stability may be affected differently with variation of
the asphaltene concentration or H/C.

3.2.2. Fit of Petroleum Emulsion Stability to Kinetic Model


The correlation is very encouraging and may serve the industry well in
predicting long-time emulsion stability of water-in-crude oil systems at conditions in
which wax issues are unimportant. In order to discover information on the kinetics of
asphaltene adsorption, critical electric fields for all petroleums with 30 % water were
determined at various times from 3 minutes to 24 hours after emulsification and the
results are displayed in Figures 11 and 12. The data points represent the measured cef

20
Kilpatrick: The Stability of Water-in-Crude and Model Oil Emulsions 21

values while the lines are the predicted trends obtained from fitting the data to the
kinetic model.
The emulsion stability data fit the model reasonably well. Some crudes showed
a significant decrease in cef at short times followed by an eventual increase to the 24
hour value. This was probably due to droplet settling rather than a weakening of the
interfaces at short times. Immediately after emulsification, all droplets were small. As
these small droplets coalesced, larger droplets formed and fell to the bottom of the
emulsion. In crude oils with high viscosities or densities close to that of water, this
settling occurred slowly, consequently, at short times larger droplets were used for cef
testing when the sample was withdrawn from the middle of the emulsion. As a result,
the water to oil ratio in the sample cell was higher and the measured cef was lower for
these emulsions. This was also observed for some petroleums which were not as
viscous and was probably due to sampling the emulsion from the wrong level. The rate
constants from the model fits for B6 and CS were not used for further analysis because
the experimental difficulties produced results that could not be fit well to the model.
The data for AB show a lag time before a cef increase, similar to that observed for 0.5
% AH asphaltenes in 40 % toluene, that may be due to its low asphaltene concentration
(0.79 %). The low bulk asphaltene concentration led to low concentration of adsorbed
asphaltenes at short times and no increase in emulsion stability. After the first 30
minutes, the concentration of adsorbed asphaltenes was high enough to provide
increased stability over the initial value, and cef increased at a rapid rate. At long
times, the cef drops, probably due to sampling error. The 24 hour cef was assumed, for
model fitting purposes, to be 0.07 kV/cm greater than the 6 hour value based on results
for most of the other petroleums.
The dependence of ka on R/A of the crude oils is shown in Figure 13. The value
of ka increases with R/A. Asphaltene aggregates are more solvated at higher values of
R/A and consequently are smaller. Smaller aggregates are more mobile and adsorb to
the water-oil interface faster because they are more influenced by interfacial forces. For
the purpose of fitting the overall cef data, resin concentration is not as important,
suggesting that the major mechanism by which resins control emulsion stability is the
modification of the driving force for adsorption of near interfacial asphaltenes.

21
Produce by The Berkely Electroni Pres, 205
22

The qualities of the fits indicate this is an adequate model to describe the
kinetics and is physically meaningful. A few important inferences can be drawn from
these fits: (1) water droplets are stabilized very rapidly by adsorption of asphaltenes to
yield an emulsion system which requires a significant electric field to induce
coalescence; (2) a longer consolidation process occurs during which the asphaltenic
film undergoes conformational changes which produce a stronger interfacial film. The
length scale for asphaltene diffusion from the bulk to the oil-water interface in these
well-mixed systems is small enough that the viscosity or density of the oil phase
probably has a very small effect on the long time stability of the emulsion and the
differences between petroleums is due to asphaltene interactions with the interface.
77
Sjöblom and coworkers investigated the effect of aging on crude oil systems
which were diluted with various amounts of solvents (0-50 % added) including
“condensate F” – a mixture of alkanes and aromatics, heptane, 50 % heptane/50 %
toluene, and toluene. For all of the systems, they observed no effect on cef for
emulsions with 20 % water tested immediately after homogenization versus those aged
one week. The stabilities ranged from 2.3 kV/cm for pure crude oil to about 0.5 kV/cm
for crude oil diluted with 50 % toluene. Both these results and the ones we have
obtained confirm that the kinetics of asphaltene-film stabilization in crude oil systems
is rapid.

3.2.3. Petroleum blends: Emulsion stability and interfacial film formation kinetics
Cef was measured for emulsions produced with AB/HO blends with 30 %
water. The results are shown in Figure 14 as cef vs. % A x H/C plot. The data do not
fall on the whole crude stability correlation, but show increased stability over the whole
range of compositions. The largest increase relative to the petroleum correlation is
found for high AB content blends. We investigated the kinetics of interfacial film
formation for each blend composition using the kinetic model in order to understand the
discrepancies between the petroleum and blend data. The fitted kinetic parameters are
listed in Table 3. The dependence of ka and kc on % A x • H/C is shown in Figure 15.
The values of ka and kc peak at 5 % HO, the point at which emulsion stability deviates
the most from the petroleum correlation. This result indicates that the positive

22
Kilpatrick: The Stability of Water-in-Crude and Model Oil Emulsions 23

deviations in emulsion stability were due to increased mobility and adsorption rate of
the asphaltene aggregates at the water-oil interface as well as more rapid consolidation.
When a small amount (1-5 %) of HO was added to AB, the H/C value of the DeA crude
did not substantially change from the value for AB, but the H/C value and
concentration of the asphaltene fraction changed significantly due to the high
asphaltene concentration in HO (14.8 %) relative to AB (0.79 %). HO asphaltenes
were better dispersed in the blend solvent and the resulting asphaltene mobility was
much higher than expected based on a simple ideal mixing assumption, producing
deviations from the petroleum correlation.
The asphaltene adsorption rates in the model emulsions with AH asphaltenes in
40 % toluene were relatively high for similar reasons as observed for those in blends
with low concentrations of HO in AB. AH and AB asphaltenes are both highly
aromatic asphaltenes (low H/C). In both the low HO blends and 40 % toluene model
oils, the H/C ratio of the solvent was higher than the original solvent the asphaltenes
were dissolved in, and significant aggregation occurred providing a high driving force
for interfacial adsorption.

4. Conclusions

Critical electric field measurement is a good technique to quantitatively gauge


emulsion stability. We have applied the critical electric field technique to model
emulsions and revealed the importance of asphaltene solvency to emulsion stability.
We have shown conclusively with these tests that both the asphaltene concentration and
solvent-asphaltene chemical mismatch are key parameters for determining emulsion
stability. The correlation of interfacial mass with cef has revealed a critical extent of
interfacial coverage, above which, emulsion stability is less affected by adsorbed
asphaltenes. These studies have shown that cef truly is a direct probe of the interface.
We have also developed a kinetic model for emulsion stabilization due to asphaltene
adsorption. With this model, we have shown stabilization to be due to asphaltene
adsorption followed by consolidation to form an interfacial film.
The ability to probe the strength of the interfacial film directly has resulted in
clear and very understandable correlations for water-in-crude oil emulsion stability with

23
Produce by The Berkely Electroni Pres, 205
24

characteristics of the system which control asphaltene interfacial adsorption. The


correlations developed have a tremendous amount of physical and chemical research
behind them. The results are physically appealing and remarkably simple. The kinetic
model has revealed information on the stabilization of petroleum emulsions.
Asphaltene aggregates that are more solvated by resins or the petroleum solvent adsorb
more rapidly to the interface. The long time emulsion stability is dictated by overall
crude oil properties, which are asphaltene-solvent chemical mismatch and asphaltene
concentration. The similar behavior of the crude and model oil systems with
asphaltenes alone, verified that asphaltene adsorption to the water-oil interface is the
primary mechanism responsible for crude oil emulsion stability. The differences when
resins are added suggest some key property for emulsion stabilization is missing in
model oils. This property is probably viscosity. In order to better investigate the
kinetics of film formation a better model oil system needs to be found which better
simulates petroleum behavior.
It is very noteworthy that a simple correlation was found for petroleum
emulsion stability considering the huge variety of petroleum properties and enormous
number of potential parameters. This indicates that despite the complexity of
petroleum, by conducting a careful, well thought out analysis, utilizing knowledge of
petroleum chemistry and model oil emulsion results, the key parameters for petroleum
emulsion stability can be identified. There are some issues that remain unaddressed. Is
it possible to extend the cef technique to lower temperatures and for crudes in which
wax solvency and precipitation plays a role? Can we extend the correlation to refinery
emulsions (specifically desalters and API separators) in which inorganic solids (iron
sulfide, iron oxide, calcium carbonate, etc.) can be present at levels of 1-5+% (w/w)?
How should the correlation account for rapidly changing solvent conditions when
crudes are blended? With crudes of higher R/A, such as SF, MI, and GM, blending
with a very stable crude with a much lower R/A, such as HO, B-4, or B-6, will likely
yield crude blends in which % resin will enter into the correlation. There are many
possible directions to follow in this research effort. Based on the results of this
investigation so far we are optimistic about future studies.

24
Kilpatrick: The Stability of Water-in-Crude and Model Oil Emulsions 25

5. Acknowledgements

The authors would like to thank Dr. P. Matthew Spiecker for performing some
of the solubility work and providing information on interfacial film masses that aided in
the calculations of film thickness. We are also grateful to the Petroleum Environmental
Research Forum for funding this work through grants 95-02 and 97-07 and to the
National Science Foundation for funding this work through research grant CTS-
9817127. JS would like to thank the Norwegian Research Council (NFR) and his oil
consortium for financial support.

6. Bibliography
1) Schramm, L. L. Petroleum Emulsions: Basic Principles; American Chemical
Society: Washington, D.C., 1992; Vol. 231, pp 1-49.
2) Schubert, H.; Armbruster, H., "Principles of Formation and Stability of Emulsions,"
International Chemical Engineering 1992, 32, 14-28.
3) Sjöblom, J.; Skodvin, T.; Holt, Ø.; Nilsen, F. P., "Colloid Chemistry and Modern
Instrumentation in Offshore Petroleum Production and Transport," Colloids and
Surfaces A: Physicochemical and Engineering Aspects 1997, 123-124, 593-607.
4) Tissot, B. P.; Welte, D. H. Petroleum Formation and Occurrence: A New Approach
to Oil and Gas Exploration; Springer-Verlag: Berlin, 1984.
5) Obah, B., "The Chemical Demulsification of Crude Oil Emulsion: Problem in a
Niger Delta Oil Terminal," Erdohl, Kohle, Erdgas Petrochemie 1988, 41, 71-74.
6) Bhardwaj, A.; Hartland, S., "Study of Demulsification of Water-in-Crude Oil
Emulsion.," Journal of Dispersion Science and Technology 1993, 14, 541-557.
7) Mohammed, R. A.; Bailey, A. I.; Luckham, P. F.; Taylor, S. E., "The Effect of
Demulsifiers On the Interfacial Rheology and Emulsion Stability of Water-in-Crude Oil
Emulsions," Colloids and Surfaces A: Physicochemical and Engineering Aspects 1994,
91, 129-139.
8) Mukherjee, S.; Kushnick, A. P. Effect of Demulsifiers on Interfacial Properties
Governing Crude Oil Demulsification; American Chemical Society: Toronto, 1988;
Vol. 33, pp 205-210.
9) Singh, B. P., "Performance of Demulsifiers: Prediction Based on Film Pressure-
Area Isotherms and Solvent Properties," Energy Sources 1994, 16, 377-385.
10) Sjoblom, J.; Soderlund, H.; Lindblad, S.; Johansen, E. J.; Skjarvo, I. M., "Water-in-
Crude Oil Emulsions from the Norwegian Continental Shelf : Part II. Chemical
Destabilization and Interfacial Tensions," Colloid Polym. Sci. 1990, 268, 389-398.
11) Wasan, D. T. Destabilization of Water-in-Oil Emulsions; Sjöblom, J., Ed.; Kluwer
Academic Publishers: Netherlands, 1992, pp 283-295.
12) Breen, P.J.; Yen, A.; Tapp, J. "Demulsification of asphaltene-stabilized emulsions-
correlations of demulsifier performance with crude oil composition.," Petroleum
Science and Technology 2003, 21, 437-447.

25
Produce by The Berkely Electroni Pres, 205
26

13) Kim, Y.-H.; Nikolov, A. D.; Wasan, D. T.; Diaz-Arauzo, H.; Shetty, C. S.,
"Demulsification of Water-in-Crude Oil Emulsions: Effects of Film Tension,
Elasticity, Diffusivity, and Interfacial Activity of Demulsifier Individual Components
and their Blends.," Journal of Dispersion Science and Technology 1996, 17, 33-53.
14) Krawczyk, M. A.; Wasan, D. T.; Shetty, C. S., "Chemical Demulsification of
Petroleum Emulsions Using Oil-Soluble Demulsifiers," Industrial and Engineering
Chemistry Research 1991, 30, 367-375.
15) Hirato, T.; Koyama, K.; Tanaka, T.; Awakura, Y.; Majima, H., "Demulsification of
Water-in-Oil Emulsion by an Electrostatic Coalescence Method," Materials
Transactions 1991, 52, 257-263.
16) Malhotra, A. K.; Wasan, D. T., "Stability of Foam and Emulsion Films: Effects of
the Drainage and Film Size on Critical Thickness of Rupture," Chemical Engineering
Communications 1986, 48, 35-56.
17) Zaki, N. N., et al., "Polyoxyethylenated Bisphenol-A for Breaking Water-in-Oil
Emulsions.," Polymers for Advanced Technologies 1996, 7, 805-808.
18) Zaki, N. N.; Abdel-Raouf, M. E.; Abdel-Azim, A.-A. A., "Propylene Oxide-
Ethylene Oxide Block Copolymers as Demulsifiers for Water-in-Oil Emulsions, I.
Effect of Molecular Weight and Hydrophilic-Lipophylic Balance on the
Demulsification Efficiency," Monatshefte fur Chemie 1996, 127, 621-629.
19) Mackay, G. D. M.; McLean, A. Y.; Betancourt, O. J.; Johnson, B. D., "The
Formation of Water-in-Oil Emulsions Subsequent to an Oil Spill," Journal of the
Institute of Petroleum 1973, 59, 164-172.
20) Strassner, J. E., "Effect of pH on Interfacial Films and Stability of Crude Oil-Water
Emulsions," Journal of Petroleum Technology 1968, 20, 303-312.
21) Dodd, C. G., "The Rheological Properties of Films at Crude Petroleum-Water
Interfaces," Journal of Physical Chemistry 1960, 64, 544-550.
22) Graham, D. E.; Stockwell, A.; Thompson, D. G. Chemical Demulsification of
Produced Crude Oil Emulsions; Ogden, P. H., Ed.; Royal Society of Chemistry:
Sunbury-on-Thames, Middlesex, U. K., 1983, pp 73-91.
23) Cratin, P. D., "A Quantitative Characterization of pH-Dependent Systems,"
Industrial and Engineering Chemistry 1969, 61, 35-45.
24) Taylor, S. E., "Resolving Crude Oil Emulsions," Chemistry and Industry 1992, 20,
770-773.
25) Siffert, B.; Bourgeois, C.; Papirer, E., "Structure and Water-Oil Emulsifying
Properties of Asphaltenes," Fuel 1984, 63, 834-837.
26) Pelet, R.; Behar, F.; Monin, J., "Resins and Asphaltenes in the Generation and
Migration of Petroleum," Organic Geochemistry 1985, 10, 481-498.
27) Yen, T. F., "The Colloidal Aspect of a Macrostructure of Petroleum Asphalt," Fuel
Science and Technology International 1992, 10, 723-733.
28) Dickie, J. P.; Yen, T. F., "Macrostructures of the Asphaltic Fractions by Various
Instrumental Methods," Analytical Chemistry 1967, 39, 1847-1852.
29) Cimino, R.; Correra, S.; Bianco, A. D.; Lockhart, T. P. Solubility and Phase
Behavior of Asphaltenes in Hydrocarbon Media; Sheu, E. Y. and Mullins, O. C., Ed.;
Plenum Press: New York, 1995, pp 97-130.
30) Speight, J. G., "Polynuclear Aromatic Systems in Petroleum," Preprints -- ACS
Division of Petroleum Chemistry 1986, 31, 818-825.

26
Kilpatrick: The Stability of Water-in-Crude and Model Oil Emulsions 27

31) Speight, J. G., "Latest Thoughts on the Molecular Nature of Petroleum


Asphaltenes," Preprints -- ACS Division of Petroleum Chemistry 1989, 34, 321-328.
32) Bestougeff, M. A.; Byramjee, R. J. Chemical Constitution of Asphaltenes; Yen, T.
F. and Chiliingarian, G. V., Ed.; Elsevier Science, 1994, pp 67-94.
33) Moschopedis, S. E.; Fryer, J. F.; Speight, J. G., "Investigation of Asphaltene
Molecular Weights," Fuel 1976, 55, 227-232.
34) Speight, J. G.; Moschopedis, S. E., "Some Observations on the Molecular "Nature"
of Petroleum Asphaltenes," Preprints -- ACS Division of Petroleum Chemistry 1979,
24, 910-923.
35) Watson, B. A.; Barteau, M. A., "Imaging of Petroleum Asphaltenes Using Scanning
Tunneling Microscopy," Indusrial & Engineering Chemistry Research 1994, 33, 2358-
2363.
36) Kawanaka, S.; Leontaritis, K. J.; Park, S. J.; Mansoori, G. A. Thermodynamic and
Colloidal Models of Asphaltene Flocculation; Borchardt, J. K. and Yen, T. F., Ed.;
American Chemical Society: Washington D C, 1989, pp 443-458.
37) Sheu, E. Y.; Storm, D. A.; Tar, M. M. D., "Asphaltenes in Polar Solvents," Journal
of Non-Crystalline Solids 1991, 131-133, 341-347.
38) Aveyard, R.; Binks, B. P.; Fletcher, P. D. I.; Lu, J. R., "The Resolution of Water-in-
Crude Oil Emulsions by the Addition of Low Molar Mass Demulsifiers," Journal of
Colloid and Interface Science 1990, 139, 128-138.
39) Lawrence, A. S. C.; Killner, W., "Emulsions of Seawater in Admiralty Fuel Oil
with Special Reference to their Demulsification," Journal of the Institute of Petroleum
1948, 34, 281.
40) Bhardwaj, A.; Hartland, S., "Dynamics of Emulsification and Demulsification of
Water in Crude Oil Emulsions," Industrial and Engineering Chemistry Research 1994,
33, 1271-1279.
41) McLean, J. D.; Kilpatrick, P. K., "Effects of Asphaltene Solvency on Stability of
Water-in-Crude Oil Emulsions," Journal of Colloid and Interface Science 1997, 189,
242-253.
42) Gelot, A.; Friesen, W.; Hamza, H. A., "Emulsification of Oil and Water in the
Presence of Finely Divided Solids and Surface-Active Agents," Colloids and Surfaces
1984, 12, 271-303.
43) Menon, V. B.; Wasan, D. T., "A Review of the Factors Affecting the Stability of
Solids-Stabilized Emulsions," Separation Science and Technology 1988, 23, 2131-
2142.
44) Yan, N.; Masliyah, J. H., "Characteriztaion and Demulsification of Solids-
Stabilized Oil-in-Water Emulsions Part 1. Partitioning of Clay Particles and
Preparation of Emuksions," Colloids and Surfaces A: Physicochemical and
Engineering Aspects 1995, 96, 229-242.
45) Yan, N.; Masliyah, J. H., "Characterization and Demulsification of Solids-
Stabilized Oil-in-Water Emuslions Part 2. Demulsification by the Addition of Fresh
Oil," Colloids and Surfaces A: Physicochemical and Engineering Aspects 1995, 96,
243-252.
46) Papirer, E.; Bourgeois, C.; Siffert, B.; Balard, H., "Chemical Nature and Water/Oil
Emulsifying Properties of Asphaltenes," Fuel 1982, 61, 732-734.

27
Produce by The Berkely Electroni Pres, 205
28

47) Shetty, C. S.; Nikolov, A. D.; Wasan, D. T., "Demulsification of Water in Oil
Emulsions Using Water Soluble Demulsifiers," Journal of Dispersion Science and
Technology 1992, 13, 121-133.
48) Ese, M.-H.; Sjöblom, J.; Førdedal, H.; Urdahl, O.; Rønningsen, H. P., "Ageing of
Interfacially Active Components and Its Effect On Emulsion Stability As Studied By
Means of High Voltage Dielectric Spectroscopy Measurements," Colloids and Surfaces
1997, 123-124, 225-232.
49) McLean, J. D.; Kilpatrick, P. K., "Effects of Asphaltene Aggregation in Model
Heptane-Toluene Mixtures on Stability of Water-in-Oil Emulsions," Journal of Colloid
and Interface Science 1997, 196, 23-34.
50) Carroll, B. J., "The Stability of Emulsions and Mechanisms of Emulsion
Breakdown," Surface and Colloid Science 1976, 9, 1-65.
51) Menon, V. B.; Wasan, D. T., "Coalescence of Water-in-Shale Oil Emulsions,"
Separation Science and Technology 1984, 19, 555-574.
52) Isaacs, E. E.; Huang, H.; Babchin, A. J.; Chow, R. S., "Electroacoustic Method for
Monitoring the Coalescence of Water-in-Oil Emulsions," Colloids and Surfaces 1990,
46, 177-192.
53) Parkinson, C.; Sherman, P., "Phase Inversion Temperature As An Accelerated
Method For Evaluating Emulsion Stability," Journal of Colloid and Interface Science
1972, 41, 328-330.
54) Stalss, F.; Böhm, R.; Kupfer, R., "Improved Demulsifier Chemistry: A Novel
Approach in the Dehydration of Crude Oil," Society of Petroleum Engineers --
Production Engineering 1991, 334-338.
55) Boyd, J.; Parkinson, C.; Sherman, P., "Factors Affecting Emulsion Stability, and
the HLB Concept," Journal of Colloid and Interface Science 1972, 41, 359-370.
56) Mikula, R. J. Emulsion Characterization; Schramm, L. L., Ed.; ACS: Washington
D.C., 1992, pp 79-130.
57) Sjöblom, J.; Fordedal, H.; Jakobsen, T.; Skodvin, T. Dielectric Spectroscopic
Characterization of Emulsions; Birdi, K. S., Ed.; CRC Press: Boca Raton, 1997, pp
217-237.
58) Skodvin, T.; Sjöblom, J.; Saeten, J. O.; Urdahl, O.; Gestblom, B., "Water-in-Crude
Oil Emulsions from the Norwegian Continental Shelf IX. A Dielectric Spectroscopic
Characterization of Authentic as Well as Model Systems," Journal of Colloid and
Interface Science 1994, 166, 43-50.
59) Skodvin, T.; Sjöblom, J., "Dielectric Spectroscopy on W/O Emulsions Under
Influence of Shear Forces," Colloid and Polymer Science 1996, 274, 754-762.
60) Førdedal, H.; Nodland, E.; Sjöblom, J.; Kvalheim, O. M., "A Multivariate Analysis
of W/O Emulsions in High External Electric Fields as Studied by Means of Dielectric
Time Domain Spectroscopy," Journal of Colloid and Interface Science 1995, 173, 396-
405.
61) Førdedal, H.; Schildberg, Y.; Sjöblom, J.; Volle, J.-L., "Crude Oil Emulsions in
High Electric Fields as Studied by Dielectric Spectroscopy. Influence of Interaction
Between Commercial and Indigenous Surfactants," Colloids and Surfaces A:
Physicochemical and Engineering Aspects 1996, 106, 33-47.
62) Bailes, P. J.; Larkai, S. K. L., "Liquid Phase Separation in Pulsed D.C. Fields,"
Transactions of the Institution of Chemical Engineers 1982, 60, 115-121.

28
Kilpatrick: The Stability of Water-in-Crude and Model Oil Emulsions 29

63) Chen, T. Y.; Mohammed, R. A.; Bailey, A. I.; Luckham, P. F.; Taylor, S. E.,
"Dewatering of Crude Oil Emulsions 4. Emulsion Resolution by the Application of an
Electric Field," Colloids and Surfaces A: Physicochemical and Engineering Aspects
1994, 83, 273-284.
64) Urdahl, O.; Williams, T. J.; Bailey, A. G.; Thew, M. T., "Electrostatic
Destabilization of Water-in-Oil Emulsions Under Conditions of Turbulent Flow,"
Transactions of the Institute of Chemical Engineers 1996, 74, 158-165.
65) Wang, L.; Yan, Z., "The Optimum Frequency in Electrically Enhanced Coalescence
of W/O Type Emulsions," Journal of Dispersion Science and Technology 1994, 15, 35-
58.
66) Waarden, M. v. d., "Stability of Emulsions of Water in Mineral Oils Containing
Asphaltenes," Kolloid Z. Z. Polymer 1958, 156, 116-122.
67) Spiecker, P. M. "The Impact of Asphaltene Chemistry and solvation on Emulsion
and Interfacial Film formation," Ph.D. Thesis, Dept. of Chem. Eng., North Carolina
State Univ., 2001.
68) Acevedo, S.; Mendez, B.; Rojas, A.; Layrisse, I.; Rivas, H., "Asphaltenes and
Resins From the Orinoco Basin," Fuel 1985, 64, 1741-1747.
69) Acevedo, S.; Escobar, G.; Gutierrez, L.; Rivas, H., "Isolation and Characterization
of Natural Surfactants from Extra Heavy Crude Oils, Asphaltenes, and Maltenes.
Interpretation of Their Interfacial Tension-pH Behaviour in Terms of Ion Pair
Formation," Fuel 1992, 71, 619-623.
70) Ali, M. F.; Saleem, M., "Asphaltenes in Saudi Arabian Heavy Crude Oil Solubility
and Molecular Weights in Hydrocarbon Solvents," Fuel Science and Technology 1988,
6, 541-556.
71) Al-Jarrah, M. M. H.; Al-Dujaili, A. H., "Characterization of Some Iraqi Asphalts
II. New Findings on the Physical Nature of Asphaltenes," Fuel Science and
Technology International 1989, 7, 69-88.
72) Ali, L. H.; Al-Ghannam, K. A.; Al-Rawi, J. M., "Chemical Structure of
Asphaltenes in Heavy Crude Oils Investigated by NMR," Fuel 1990, 69, 519-521.
73) Calemma, V.; Iwanski, P.; Nali, M.; Scotti, R.; Montanari, L., "Structural
Characterization of Asphaltenes of Different Origins," Energy & Fuels 1995, 9, 225-
230.
74) Cyr, N.; McIntyre, D. D.; Toth, G.; Strausz, O. P., "Hydrocarbon Structural Group
Analysis of Athabasca Asphaltene and its G.P.C. Fractions by C-13 N.M.R.," Fuel
1987, 66, 1709-1714.
75) McKay, J. F.; Amend, P. J.; Cogswell, T. E.; Harnsberger, P. M.; Erickson, R. B.;
Latham, D. R. Petroleum Asphaltenes: Chemistry and Composition; Uden, P. C.,
Siggia, S. and Jensen, H. B., Ed.; American Chemical Society: Washington D.C., 1978;
Vol. 170, pp 128-142.
76) Storm, d. A.; DeCanio, S. J.; DeTar, M. M.; Nero, V. P., "Upper Bound on Number
Average Molecular Weight of Asphaltenes," Fuel 1990, 69, 735-738.
77) Mouraille, O.; Skodvin, T.; Sjoblom, J.; Peytavy, J.-l., "Stability of Water-in-Crude
Oil Emulsions: Role Played by the State of Solvation of Asphaltenes and by Waxes," J.
Dispersion Science and Technology 1998, 19, 339-367.
78) Jeribi, M.; Almir-Assad, B.; Langevin, D.; Henaut, I.; Argillier, J. F. “Adsorption
kinetics of asphaltenes at liquid interfaces.” J. Colloid and Interface Science 2002,
256, 268-272

29
Produce by The Berkely Electroni Pres, 205
30

Table 1: Summary of Crude Properties


Crude %Aa %Rb R/A H/C H/C H/C H/C H/C Density Viscosity Kin Visc
Wholec Ac Rc DeAd (DeA - (g/mL) @100F @100F
A) @60Fe (cP)f (cSt)g
h
Arab Hvy II 6.68 7.46 1.12 1.683 1.080 1.372 1.725 0.645 0.946 33.8 35.7
(AH)
Arab 0.79 3.24 4.10 1.814 1.020 1.349 1.820 0.800 0.838 4.39 5.24
Berri (AB)
Alaska North 3.38 8.72 2.58 1.710 1.057 1.408 1.732 0.675 0.889 12.8 14.4
Slope (ANS)
San Joaquin 4.56 19.4 4.25 1.518 1.170 1.38 1.535 0.365 0.979 1390 1420
Valley (SJV)
B-4 13.6 12.2 0.90 1.593i 1.222 1.514 1.667 0.445 0.935 2310 2470

B-6 13.1 12 0.92 1.553i 1.224 1.536 1.618 0.394 0.935 2030 2170

Alba 1.64 10.1 6.17 1.651 1.144 1.433 1.659 0.515 0.940 136 145

Malu Isan 0.18 4.86 27.0 1.913 1.333 1.499 1.914 0.581 0.845 38.2 45.2
(MI)
Sour Maya 11.5 11 0.95 1.615 1.087 1.410 1.682 0.595 0.919 75 81.6
(SM)
Canadon 7.5 8.94 1.19 1.680 1.028 1.376 1.734 0.706 0.903 70 77.5
Seco (CS)
Statfjord 0.09 4.02 42.6 1.844 1.289 1.412 1.845 0.556 0.833 3.14 3.77
(SF)
Gulf of 0.31 5.02 16.2 1.782 1.117 1.374 1.784 0.667 0.870 7.11 8.17
Mexico (GM)
South China 3.46 6.05 1.75 1.867 1.353 1.386 1.886 0.533 0.858 --- ---
Sea(SCS)
Thums 1 5.09 18.7 3.67 1.696 1.153 1.455 1.726 0.573 0.952 152 160
(TH1)
Thums 2 3.31 12.5 3.77 1.690i 1.178 1.442 1.718 0.50 1.01h 656 650
(TH2)
Hondo (HO) 14.8 20.5 1.39 1.667 1.248 1.508 1.738 0.490 0.938 363 387

a = n-heptane precipitation
b = sequential elution chromatography
c = combustion and GC analysis (Galbraith Laboratories)
d = calculated from combustion and GC analyses of whole crude, A, and R
{H,C = whole(H,C)-%A*[A(H,C)])}
e = from assays (supplier of crude)
f = rheology (stress rheometer, couette geometry)
g = dynamic viscosity/density
h = pycnometry
i = corrected numbers based on assumption 100-%C-%H=%O from water in crude sample
(for high water content crudes)
Kilpatrick: The Stability of Water-in-Crude and Model Oil Emulsions 31

Table 2: Model Parameters for Crude Oil Emulsions ( 1=0.2, 2=0.8)


Crude Oil ka (cm3min-1g-1) kc (x103 min-1)
AB 2.5 6
AH 0.1 3
AL 1.2 3
ANS 3.9 2
B4 1.2 4
B6 0.3 5
CS 4.8 9
HO 0.1 16
SCS 10 2
SJV 2.6 2
SM 0.8 4
TH1 0.3 4

Produce by The Berkely Electroni Pres, 205


32

Table 3: Model Parameters for Crude Oil Blend Emulsions ( 1=0.2, 2=0.8)
%HO in AB/HO ka (cm3min-1g-1) kc (x103 min-1)
0.0 2.5 6
1.0 4.8 7
5.0 14 67
11.4 0.2 20
18.0 0.6 23
25.0 1.0 4
50.0 1.2 6
75.0 1.9 4
100 0.1 16
Kilpatrick: The Stability of Water-in-Crude and Model Oil Emulsions 33

Figure Captions

Figure 1: Asphaltene interactions in crude oil. a) formation of an asphaltene aggregate solvated with resin
molecules to shield polar functionality from crude oil medium, b) adsorption of asphaltene aggregates to a
water-oil interface and crosslinking through H-bond and dipole-dipole interactions to form a viscoelastic
film.

Figure 2: Schematic of water-in-oil emulsion sample in the cef cell. Initially the disperse water droplets are
evenly distributed throughout the sample. Droplets aggregate in chains parallel to the electric field as the
field strength is increased. At a certain field strength (cef), a continuous water bridge spans the electrode
gap.

Figure 3: Side-view of the sample cell used for cef measurements

Figure 4: Solvent effect on the cef of model emulsions (60 % water) stabilized with AH asphaltenes and
asphaltene solubility. Again, maximum emulsion stability coincides with the solubility limit of the
asphaltenes. Increased solubility of AH relative to HO asphaltenes translates to higher cef values.

Figure 5: Solvent effect on the cef of model emulsions (30 % water) with AH asphaltenes. There is a critical
asphaltene concentration, above which, emulsion stability does not change much with increased asphaltene
concentration.

Figure 6: Droplet diameters in AH asphaltene stabilized model emulsions (30 % water).

Figure 7: Adsorbed asphaltene mass versus asphaltene concentration in model oil emulsions. The change in
interfacial mass with bulk asphaltene concentration is similar at all soluble conditions. For HO asphaltenes,
the solubility limit was reached at 2 % and the interfacial film mass/area remained constant with increased
concentration.

Figure 8: Model emulsion stability as a function of calculated interfacial film mass/area and interfacial
thickness. Emulsion stability increases with interfacial thickness up until a critical extent of interfacial
coverage.

Figure 9: Variation in AH asphaltene-stabilized model oil emulsion stability with time. Emulsion
stabilization occurs in a two step process: rapid interfacial adsorption, followed by long term film
consolidation.

Figure 10: Petroleum emulsion stability (30 % water) as a function of the % A x • H/C.

Figure 11: Variation in petroleum emulsion stability (30 % water) with time.

Figure 12: Variation in petroleum emulsion stability (30 % water) with time.

Figure 13: Dependence of the asphaltene adsorption rate constant (ka) on R/A in petroleum emulsions (30%
water). Increased R/A results in higher adsorption rates due to larger aggregate mobility.

Produce by The Berkely Electroni Pres, 205


34

Figure 14: Emulsion stability of AB/HO petroleum blends (30 % water) as a function of A x H/C. Blend
emulsion stabilities show a positive deviation over the entire range of compositions.

Figure 15: Dependence of rate constants (ka and kc) on A x H/C in AB/HO blend emulsions. The
maximum value of the rate constants corresponds to the blend that shows the largest discrepancy in stability
relative to the petroleum correlation.
a.
aromatic core

Figure 1
polar functional group

polar functional group


aliphatic side-chain

Asphaltene "molecule" Resin "molecule" Resin-Solvated asphaltene aggregate

oil
Kilpatrick: The Stability of Water-in-Crude and Model Oil Emulsions

b.

Produce by The Berkely Electroni Pres, 205


oil-water interface
water
35
36

+ Screws to seal chamber

hole drilled copper


through to electrodes
inside chamber

aluminum
Mylar casing
spacer

insulating
water material
jacket

Figure 2
Kilpatrick: The Stability of Water-in-Crude and Model Oil Emulsions 37

+ +
V

- V
-
0 kV/cm droplets form short chains parallel
to electric field

+ +
V

- -
droplet chains lengthen and bridge critical electric field - water bridges
entire gap electrodes, conductivity jumps
Figure 3

Produce by The Berkely Electroni Pres, 205


38

60
0.40 % AH Asphaltene
2.0
0.35 3.0 50

% Precipitated Asphaltenes
5.0
7.0
0.30
40
0.25
cef (kV/cm)

30
0.20

0.15 20

0.10
10
0.05

0.00 0
20 25 30 35 40 45 50 55 60
% Toluene in Heptol

Figure 4
Kilpatrick: The Stability of Water-in-Crude and Model Oil Emulsions 39

% AH Asphaltene
0.25
0.5
1.5 1.0
2.0
3.0
cef (kV/cm)

1.0

0.5

0.0

32 36 40 44 48 52 56 60

% Toluene in Heptol

Figure 5
Produce by The Berkely Electroni Pres, 205
40

12 % Toluene
Droplet Mean Sauter Diameter (µm)

45
50
10 55

0.0 0.4 0.8 1.2 1.6 2.0 2.4 2.8 3.2

% AH Asphaltene
Figure 6
Kilpatrick: The Stability of Water-in-Crude and Model Oil Emulsions 41

40
Calculated Interfacial Mass/Area

35

30
(mg/m 2 )

25

20
A type, % toluene, % water
15 AH, 45, 30 slope = 10
AH, 50, 30 slope = 13
10
AH, 55, 30 slope = 9.4

5 HO, 50, 60 slope = 14

1 2 3 4 5
% Asphaltene

Figure 7
Produce by The Berkely Electroni Pres, 205
42

Calculated Interfacial Film Thickness (# molecules)


3.0 5.9 8.9 11.8 14.8 17.7

Critical
1.6 interfacial
coverage
1.4

1.2
cef (kV/cm)

1.0

0.8 A type, % toluene, % water


AH, 40, 30
0.6 AH , 45, 30
AH, 50, 30
0.4
AH, 55, 30

0.2 HO, 40, 60


HO, 50, 60
0.0
1 2 3 4 5 6
2
Calculated Interfacial Mass/Area (mg/m )
Figure 8
Kilpatrick: The Stability of Water-in-Crude and Model Oil Emulsions 43

1.4

1.2

1.0
cef (kV/cm)

0.8

0.6
1 % AH, 40 % tol
0.4 0.5 % AH, 40 % tol
1 % AH, 50% tol

0.2

0 200 400 600 800 1000 1200 1400


Time (minutes)

Figure 9
Produce by The Berkely Electroni Pres, 205
44

3 .0
HO

2 .5 SM

2 .0
cef(KV/cm)

AH
B6 B4
1 .5 CS

AL SJV
1 .0 TH1
ANS
0 .5
AB
0 .0
0 1 2 3 4 5 6 7 8
A x H /C

Figure 10
Kilpatrick: The Stability of Water-in-Crude and Model Oil Emulsions 45

3.0

2.5

2.0
cef (kV/cm)

1.5

HO SCS
1.0
CS B4
SM B6
0.5

0.0
0 200 400 600 800 1000 1200 1400
Time (min)

Figure 11

Produce by The Berkely Electroni Pres, 205


46

1.8

1.6

1.4

1.2
cef (kV/cm)

1.0

0.8
AH AL
0.6
TH 1 ANS
0.4 SJ V AB

0.2
0 200 400 600 800 100 0 120 0 140 0
T im e (m in )

Figure 12
Kilpatrick: The Stability of Water-in-Crude and Model Oil Emulsions 47

ka = 0.13 + 0.30*R/A; R = 0.79


2.0
ka

1.5
ka (cm 3 min -1 g -1 )

1.0

0.5

0.0
0.8 1.6 2.4 3.2 4 4.8 5.6 6.4
R/A

Figure 13
Produce by The Berkely Electroni Pres, 205
48

3.0
AB/HO Blend
Petroleum Correlation
2.5

2.0
cef (kV/cm)

1.5

1.0

0.5

0.0
0 1 2 3 4 5 6 7 8
A x delta H/C
Figure 14
Kilpatrick: The Stability of Water-in-Crude and Model Oil Emulsions 49

0.08
12 k k
a c
0.07

10 0.06
ka (cm 3 min -1 g -1 )

8 0.05

kc (min -1 )
0.04
6
0.03
4
0.02
2
0.01

0 0
0 1 2 3 4 5 6 7 8
A * delta H/C
Figure 15

Produce by The Berkely Electroni Pres, 205

You might also like